proof
Definable convolution and idempotent Keisler measures III. Generic stability, generic transitivity, and revised Newelski’s conjecture
Abstract.
We study idempotent measures and the structure of the convolution semigroups of measures over definable groups.
We isolate the property of generic transitivity and demonstrate that it is sufficient (and necessary) to develop stable group theory localizing on a generically stable type, including invariant stratified ranks and connected components. We establish generic transitivity of generically stable idempotent types in important new cases, including abelian groups in arbitrary theories and arbitrary groups in rosy theories, and characterize them as generics of connected type-definable subgroups.
Using tools from Keisler’s randomization theory, we generalize some of these results from types to generically stable Keisler measures, and classify idempotent generically stable measures in abelian groups as (unique) translation-invariant measures on type-definable fsg subgroups. This provides a partial definable counterpart to the classical work of Rudin, Cohen and Pym for locally compact topological groups.
Finally, we provide an explicit construction of a minimal left ideal in the convolution semigroup of measures for an arbitrary countable NIP group, from a minimal left ideal in the corresponding semigroup on types and a canonical measure constructed on its ideal subgroup. In order to achieve it, we in particular prove the revised Ellis group conjecture of Newelski for countable NIP groups.
1. Introduction
We study idempotent measures and the structure of the convolution semigroups on measures in definable groups, as well as some related questions about topological dynamics of definable actions (continuing [CG22, CG23]).
We first recall the classical setting. If is a locally compact group and is the space of regular Borel probability measures on , one extends group multiplication on to convolution on : if and is a Borel subset of , then
A measure is idempotent if . A classical line of work established a correspondence between compact subgroups of and idempotent measures in , in progressively broader contexts [KI40, Wen54, Coh60, Rud59, Gli59] culminating in the following:
Fact 1.1.
[Pym62, Theorem A.4.1] Let be a locally compact group and . Then the following are equivalent:
-
(1)
is idempotent.
-
(2)
The support of is a compact subgroup of and is the normalized Haar measure on .
We are interested in a counterpart of this phenomenon in the definable category. In the same way as e.g. algebraic or Lie groups are important in algebraic or differential geometry, the understanding of groups definable in a given first-order structure (or in certain classes of first-order structures) is important for model theory and its applications. The class of stable groups is at the core of model theory, and the corresponding theory was developed in the 1970s-1980s borrowing many ideas from the study of algebraic groups over algebraically closed fields (with corresponding notions of connected components, stabilizers, generics, etc., see [Poi01]). More recently, many of the ideas of stable group theory were extended to the class of NIP groups, which contains both stable groups and groups definable in -minimal structures or over the -adics. This led to multiple applications, e.g. a resolution of Pillay’s conjecture for compact o-minimal groups [HPP08] or Hrushovski’s work on approximate subgroups [Hru12], and brought to light the importance of the study of invariant measures on definable subsets of the group (see e.g. [Che18] for a short survey), as well as the methods of topological dynamics (introduced into the picture starting with Newelski [New09]). In particular, deep connections with tame dynamical systems as studied by Glasner, Megrelishvili and others (see e.g. [Gla07, Gla18]) have emerged, and play an important role in the current paper.
More precisely, we now let be a group definable in some structure (i.e. both the underlying set and multiplication are definable by formulas with parameters in ), it comes equipped with a collection of definable subsets of cartesian powers of closed under Boolean combinations, projection and Cartesian products (but does not carry topology or any additional structure a priori). We let be a “non-standard” elementary extension of , and we let denote the group obtained by evaluating in the formulas used to define in (which in the case of an algebraic group corresponds to working in the universal domain, in the sense of Weil). So e.g. if we start with the field of reals, and its additive group, then is the additive group of a large real closed field extending which now contains infinitesimals — i.e., it satisfies a saturation condition: every small finitely consistent family of definable sets has non-empty intersection. It is classical in topological dynamics to consider the action of a discrete group on the compact space of ultrafilters on , or more precisely ultrafilters on the Boolean algebra of all subsets of . In the definable setting, given a definable group , we let denote the space of ultrafilters on the Boolean algebra of definable subsets of , hence the space (called the space of types of ) is a “tame” analogue of the Stone-Čech compactification of the discrete group . Then acts on by homeomorphisms, and the same construction applies to giving the space of ultrafilters on the definable subsets of . Similarly, we let denote the space of finitely additive probability measures on the Boolean algebra of definable subsets of (and for ), it is affinely homeomorphic to the space of all regular -additive Borel probability measures on (respectively on ), with weak∗-topology. The set embeds into as realized types, and we let denote its closure (model theoretically, this corresponds to the set of global types in that are finitely satisfiable in ). Similarly, we let denote the closed convex hull of in (this is the set of global Keisler measures on finitely satisfiable in , equivalently the set of measures supported on — see [CG22, Proposition 2.11]). Similarly to the classical case, in many situations (including the ones discussed in the introduction) we have a well-defined convolution operation on (see Definition 3.35 and the discussion around it).
In this context, generalizing a classical fact about idempotent types in stable groups [New91], we have the following definable counterpart of Fact 1.1 for stable groups:
Fact 1.2.
[CG22, Theorem 5.8] Let be a (type-)definable group in a stable structure and a measure. Then is idempotent if and only if is the unique left-invariant (and the unique right-invariant) measure on a type-definable subgroup of (namely, the left-/right-stabilizer of ).
This suggests a remarkable analogy between the topological and definable settings, even though Fact 1.2 is proved using rather different methods.
In the first part of the paper (Sections 2 and 3), we study generalizations of Fact 1.2 beyond the limited context of stable groups (we note that this correspondence fails in general NIP groups without an appropriate tameness assumption on the idempotent measure [CG23, Example 4.5]). An important class of groups arising in the work on Pillay’s conjectures is that of groups with finitely satisfiable generics, or fsg groups in short [HPP08]. It contains stable groups, as well as (definably) compact groups in -minimal structures, and provides a natural counterpart to the role that compact groups play in Fact 1.2. By a well-known characterization in the NIP context (see e.g. [Sim15, Proposition 8.33]), these are precisely the groups that admit a (unique) translation-invariant measure on their definable subsets which is moreover generically stable: a sufficiently long random sample of elements from the group uniformly approximates the measure of all sets in a definable family of subsets with high probability (i.e. is a frequency interpretation measure, or fim measure, satisfying a uniform version of the weak law of large numbers — this notion is motivated by Vapnik-Chervonenkis theory, and serves as a correct generalization of generic stability for measures outside of NIP, by analogy with generically stable types in the sense of [PT11]); see Section 3.3). An analog of Fact 1.2 would thus amount to demonstrating that such subgroups are the only source of idempotent generically stable measures (see Problem 3.41).
First, in Section 2 we focus on the case of idempotent types in (i.e. -measures, equivalently ultrafilters on the Boolean algebra of definable subsets of ). After reviewing some preliminaries on generically stable types (Sections 2.1 and 2.4), we revise the case of groups in stable structures (Section 2.7), and then resolve the question in several important cases:
Theorem 1.3.
Assume is generically stable and idempotent, and one of the following holds:
- (1)
- (2)
- (3)
- (4)
Then is the unique left-/right-invariant type on a type-definable subgroup of (namely, the left-/right-stabilizer of ).
The proof proceeds by establishing the crucial property of generic transitivity (see Section 2.4) for idempotent generically stable types in these cases, namely that if , then (using local weight arguments in case (2), and the appropriate version of the theory of stratified ranks in the other cases). The question whether every generically stable idempotent type is generically transitive remains open, even for NIP groups (see Problem 2.15 and discussion in Section 2.4).
We further investigate generic transitivity, and demonstrate that it is a sufficient and necessary condition for developing some crucial results of stable group theory localizing on a generically stable type (some other elements of stable group theory for generically stable types were considered in [Wan22]). Sometimes we use a slightly stronger technical assumption that is generically stable for all , which always holds in NIP structures. In Section 2.11, working in an arbitrary theory, we define an analog of the stratified rank in stable theories restricting to subsets of defined using parameters from a Morley sequence in a generically stable type , demonstrate finiteness of this rank (Lemma 2.39) and show that this rank is left invariant (under multiplication by realizations of ) if and only if is generically transitive (Proposition 2.41). A fundamental theorem of Hrushovski [Hru90] demonstrates that in a stable theory, every type definable group (i.e. an intersection of definable sets that happens to be a group) is in fact an intersection of definable groups. The main result of Section 2.12 is an analog for generically transitive types:
Theorem 1.4 (Proposition 2.44).
If is type-definable and is generically stable, idempotent and generically transitive, then its stabilizer is an intersection of -definable groups.
Finally, in Section 2.13 we establish a chain condition for groups type-definable using parameters from a Morley sequence of a generically stable type , implying that there is a smallest group of this form — and it is equal to the stabilizer of when is generically transitive (see Lemma 2.46 and Proposition 2.50 for the precise statement).
In Section 3, we generalize some of these results from types (i.e. -measures) to general measures, in arbitrary structures. Our main result is a definable counterpart of Fact 1.1 for arbitrary abelian group:
Theorem 1.5.
(Theorem 3.45) Let be an abelian group and a generically stable measure. Then is idempotent if and only if is the unique left-invariant (and the unique right-invariant) measure on a type-definable subgroup of (namely, its stabilizer).
Groups as in Theorem 1.5, i.e. supporting an invariant generically stable measure, are called fim groups (see Section 3.6), and in the NIP case correspond precisely to fsg groups (but this is potentially a stronger condition in general). Our proof of Theorem 1.5 relies on several ingredients of independent interest. First, we develop some theory of fim groups, generalizing from fsg groups in NIP structures (Section 3.6). Then, in Section 3.4, we provide a characterization of generically stable measures of independent interest extending [CGH23], demonstrating that the usual property — any Morley sequence determines the measure of arbitrary formulas by averaging along it — holds even when the parameters of these formulas are allowed to be “random”. More precisely:
Theorem 1.6 (Theorem 3.13).
Let be fim, arbitrary, a formula, , and let . Suppose that is arbitrary such that and . Then
Moreover for every there exists so that for any as above, we have for all but many .
This is new even in the NIP case, and relies on the use of Keisler randomization theory. Namely, we use the correspondence between measures in and types in its randomization, viewed as a structure in continuous logic, that was introduced in [BY09] (and studied further in [CGH23]). It allows us to imitate in Section 3.8 the bounded local weight argument from Section 2.5 in a purely measure theoretic context, using an adapted version of generic transitivity (see Section 3.7) and arguments with pushforwards.
Problem 2.15 on whether every generically stable idempotent type is generically transitive generalizes to measures (see Problem 3.41). In Section 3.9, we distinguish a weaker property of a measure than being generically transitive, which we call support transitivity. It leads to a weaker conjecture saying that every generically stable idempotent measure is support transitive (see Problem 3.48). While this conjecture is open, it trivially holds for idempotent types, and so one can expect that if the techniques used for types in Sections 2.7–2.10 could be adapted to measures, they would rather not prove the main conjecture that every generically stable idempotent measure is generically transitive, but reduce it to the above weakening. The idea is to pass to the randomization of the structure in question, and if this randomization happens to have a well-behaved stratified rank, then apply a continuous logic version of the arguments from Sections 2.7–2.10. In Section 3.10, we illustrate how it works for stable theories (recall that stability is preserved under randomization).
In Section 4, instead of considering an individual (idempotent) measure, we study the structure of the (left-continuous compact Hausdorff) semigroup of measures on a definable NIP group under convolution, through the lens of Ellis theory. It was demonstrated in [CG23, Theorem 5.1] that the ideal (or Ellis) subgroup of any minimal left ideal is always trivial, and that when is definably amenable (i.e. admits a left-invariant finitely additive probability measure on its definable subsets), then any minimal left ideal itself is trivial, but has infinitely many extreme points when is not definably amenable. In the general, non-definably amenable case, a description of a minimal left ideal in was obtained under some additional strong assumptions (see [CG23, Theorem 6.11] and discussion at the end of Section 4.2). Here we obtain a description of a minimal left ideal of for an arbitrary countable NIP group:
Theorem 1.7 (Corollary 5.18).
Assume that is group definable in a countable NIP structure , and let be a minimal left ideal in and an idempotent. Then the ideal group carries a canonical invariant Keisler measure (see Proposition 4.15 for the definition), and is a minimal left ideal of , where denotes the space of all measures supported on , and is an idempotent in .
Theorem 1.7 is deduced using a combination of two results of independent interest that we now discuss.
An important general fact from topological dynamics is that the ideal group of the Ellis semigroup of any flow is always a compact (not necessarily Hausdorff) semi-topological group (i.e. multiplication is separately continuous) with respect to a canonical topology, the so called -topology (which is weaker than the induced topology from the Ellis semigroup). This topology was defined by Ellis and has played an essential role in the most important structural results in abstract topological dynamics, starting from the Furstenberg structure theorem for minimal distal flows (e.g. see [Aus88]) and ending with a recent theorem of Glasner on the structure of tame, metrizable, minimal flows [Gla18]. In model theory, the -topology on the ideal groups played a key role in applications to the quotients of definable groups by their model-theoretic connected components ([KP17]) and to Lascar strong types and quotients by arbitrary bounded invariant equivalence relations [KPR18, KR20]. It also partly motivated the work of Hrushovski on definability patterns structures with spectacular applications to additive combinatorics [Hru19, Hru20]. In [KP23], the -topology was used to give a shorter and simpler proof of the main result of [Hru20]. As the key result of Section 4 we demonstrate the following:
Theorem 1.8.
(Lemma 4.14) Assume that is group definable in an arbitrary NIP structure , and that the -topology on the ideal group of the -flow is Hausdorff. Then for any clopen subset of , the subset of is constructible, and so Borel, in the -topology.
It follows that when the -topology on is Hausdorff, the ideal group is a compact topological group (Corollary 4.9), so we have the unique (normalized) Haar measure on Borel subsets, and by Theorem 1.8 it induces the aforementioned Keisler measure via (Proposition 4.15). This reduces the question to understanding when the -topology is Hausdorff — which is precisely the revised version of Newelski’s conjecture (see [KP23, Conjecture 5.3]).
The Ellis group conjecture of Newelski [New09] is an important prediction in the study of NIP groups connecting a canonical model-theoretic quotient of a definable group and a dynamical invariant of its natural action on . Let be a group definable in a structure , and let be the ideal (Ellis) group of the -flow . We let be the smallest type-definable over subgroup of of bounded index. The Ellis group conjecture of Newelski says that the group epimorphism given by , for a type and a realization of , is an isomorphism under suitable tameness assumptions on the ambient theory. This conjecture was established for definably amenable groups definable in -minimal structures in [CPS14], and for definably amenable groups in arbitrary NIP structures in [CS18]. On the other hand, it was refuted for in [GPP15]. Newelski’s epimorphism was refined in [KP17] to a sequence of epimorphisms
where is the smallest bounded index subgroup of invariant under the action of , and is the subgroup of given by the intersection of the -closures of all -neighborhoods of . With this refinement, Newelski’s conjecture fails when , in which case also is not an isomorphism. The first such example, the universal cover of , was found in [CP12], and further examples were given in [GK15]. On the other hand, no examples of NIP groups with non-trivial (equivalently, with not Hausdorff in the -topology) were known. This motivated the following weakening of Newelski’s conjecture:
Conjecture 1.9.
[KP23, Conjecture 5.3] If is NIP, then the -topology on is Hausdorff.
It clearly holds whenever is finite. It is also known to hold for definably amenable groups in NIP theories, as the full Newelski’s conjecture holds in this context. Besides those two general situations, it was confirmed only for (we refer to [KP23, Section 5] for a proof and a more detailed discussion).
In Section 5 we establish the revised Newelski’s conjecture for countable NIP groups:
Theorem 1.10 (Theorem 5.17).
The revised Newelski’s conjecture holds when is a definable group in a countable NIP structure.
This relies on the fundamental theorem of Glasner describing the structure of minimal tame metrizable flows [Gla18] and a presentation of the -flow as the inverse limit of all (the Stone space of the -algebra generated by the finite set ), where ranges over all finite collections of externally definable subsets of .
2. Idempotent generically stable types
Throughout this section, we let be a complete theory, , and a monster model.
2.1. Generically stable types
We will need some basic facts about generically stable types. As usual, given a global type (automorphism-)invariant over a small set and a bigger monster model with respect to , we let be the unique extension of to a type in which is invariant over . Given another -invariant type , we define the -invariant type via for some/any in such that and . Given an arbitrary linear order , a sequence in is a Morley sequence in over if for all . Then the sequence is indiscernible over , and for any other Morley sequence in over we have . We can then define a global -invariant type as
Equivalently, for a Morley sequence in over , where be a monster model with respect to and is the unique extension of to a type in which is invariant over . For any (viewed as an ordinal) we have .
We do not assume NIP unless explicitly stated, and use the standard definition from [PT11]: a global type is generically stable if it is -invariant for some small , and for any ordinal (or just for ), a Morley sequence in over and formula , the set is either finite or co-finite.
In the following fact, items (1)–(4) can be found in [Cas11, Section 9], (5) in [GOU13, Theorem 2.4], (6) is an immediate consequence of stationarity in (2), and (7) is [PT11, Proposition 2.1]. We let denote forking independence. We will freely use some of the basic properties of forking in arbitrary theories, e.g. extension and left transitivity, see [CK12, Section 2] for a reference.
Fact 2.1.
Let be a generically stable type, invariant over a small subset . Then the following hold.
-
(1)
Every realization of is a totally indiscernible sequence over .
-
(2)
The type is the unique global non-forking extension of .
-
(3)
For any and in such that does not fork over , we have (this holds for any when is an extension base, e.g. when ).
-
(4)
In particular, if , then .
-
(5)
If is an extension base, and (where ) forks/divides over , then is inconsistent.
-
(6)
Let and let be arbitrary small tuples in . If and , then ;
-
(7)
is definable over .
Remark 2.2.
By Fact 2.1(1), for a global generically stable type invariant over , we will also be using (without mentioning) an equivalent definition of a Morley sequence with the reversed order. That is, given a linear order , we might say that is a Morley sequence in over if for all , where is with the reversed ordering. So e.g. we will refer to with as a Morley sequence in over , and write . We will frequently use this without further mention.
Fact 2.3.
[DK12, Proposition 1.2] Let be a generically stable type, invariant over a small set of parameters . Suppose that , is an element of such that , and . Then is generically stable over .
2.2. Setting
Let be an -type-definable group (in the sort of corresponding to the tuple of variables ) and . By we mean an -definable function (from to ) whose restriction to is the group operation on . Similarly, by -1 we mean an -definable function (from to ) whose restriction to is the inverse in . By compactness, we can fix a formula implied by the partial type such that: is defined and associative on ; and for all ; if then for all (but is not necessarily closed under ). As usual, for , denotes the set of types concentrated on , i.e. such that .
Given global -invariant types (), we define via for all . Together with this operation, the set of all global -invariant types in forms a left-continuous semigroup. We say that an invariant type is idempotent if .
2.3. Generically stable groups
Definition 2.4.
[PT11, Definition 2.1] A type-definable group is generically stable if there is a generically stable which is left -invariant (we might use “-invariant” and “-invariant” interchangeably when talking about global types).
Fact 2.5.
[PT11, Lemma 2.1] Suppose that is a generically stable type-definable group in an arbitrary theory, witnessed by a generically stable type . Then we have:
-
(1)
is the unique left -invariant and also the unique right -invariant type;
-
(2)
(where for some/any in a bigger monster model ).
By Fact 2.4 and its symmetric version, we get:
Corollary 2.6.
A type-definable group is generically stable if and only if there is a generically stable which is right -invariant.
2.4. Idempotent generically stable types and generic transitivity: main conjecture
Let be a generically stable type over . The following is standard:
Proposition 2.7.
The left stabilizer of is an intersection of relatively -definable subgroups of ; in particular, it is -type-definable.
Proof.
By compactness, we can find a formula implied by such that .
Since is generically stable over , it is definable over (by Fact 2.1(7)). For any formula which implies (and is an arbitrary tuple of variables) let
By the definability of over , is definable over . Then
where , is an -relatively definable subgroup of (the fact that it is a subgroup of follows from the observation that it is the stabilizer of the set under the left action of on given by , which uses the choice of ). By the choice of , we get that is an intersection of relatively -definable subgroups of . ∎
Example 2.8.
Let be an arbitrary type-definable subgroup of which is generically stable, witnessed by a generically stable left or right -invariant type . Then is obviously idempotent.
Our central question in the case of types is whether this is the only source of generically stable idempotent types:
Definition 2.9.
For the rest of the section, we let and be the left and the right stabilizer of , respectively.
Problem 2.10.
Assume that is generically stable and idempotent, and let or . Is it true that and the group is generically stable (Definition 2.4)?
We will note now that the second part follows from the first, and that the left and the right versions of the problem are equivalent.
Remark 2.11.
Assume is generically stable and , where or . Then:
-
(1)
is a generically stable group, witnessed by (hence is both the unique left-invariant and the unique right-invariant type of , by Fact 2.5);
-
(2)
is the smallest among all type-definable subgroups of with ;
-
(3)
is both the left and the right stabilizer of in .
Proof.
We will do the case of . The proof for is symmetric.
(1) As then is a left -invariant generically stable type in .
(2) is type-definable by Proposition 2.7. For any type-definable with , the group is type-definable with and . If the index is , we have some with , and , so — a contradiction. So , and .
(3) is type-definable by a symmetric argument as in Proposition 2.7. By (1), is right -invariant, so we have , and so . By the right version of (2) (which is obtained by a symmetric argument as in (2)), we conclude that . ∎
By Remark 2.11, we see that if and only if , so the left and the right versions of Problem 2.10 are indeed equivalent.
Let be a generically stable type. Let be a monster model with respect to and (the unique extension of to a type in which is invariant over ). Then still is definable over and . Pick an arbitrary .
Let be a generically stable type, and let be a monster model with respect to and (the unique extension of to a type in which is invariant over ). Then still is definable over and , . Pick an arbitrary .
Remark 2.12.
The following conditions are equivalent for a generically stable type :
-
(1)
;
-
(2)
;
-
(3)
for any (equivalently, some) , ;
-
(4)
;
-
(5)
;
-
(6)
for any (equivalently, some) , .
Proof.
. This is because , where the inclusion follows by (1).
. Pick . Then and by (2). So .
. Suppose (1) fails. Then . So for any we have that does not realize . Using Fact 2.1, this implies that does not realize , because otherwise, since (which follows from and left transitivity of forking), is the unique non-forking extension of which is exactly , a contradiction. Thus, does not realize which contradicts (3).
are obtained by symmetric arguments to the above ones.
follows from Remark 2.11. ∎
Definition 2.13.
We will say that a generically stable type is generically transitive if it satisfies any of the equivalent conditions in Remark 2.12.
Remark 2.14.
Problem 2.15.
Assume that is generically stable and idempotent. Is it then generically transitive?
In the following sections, we will provide a positive solution under some additional assumptions on the group. We will also see that elements of stable group theory (stratified rank, connected components, etc.) can be developed in an arbitrary theory localizing on a generically stable type , if and only said type is generically transitive.
We conclude this section with a couple of additional observations.
Lemma 2.16.
If is generically stable over and idempotent, then the type for some/any is also generically stable over and idempotent.
Proof.
Assume the hypothesis. Then is generically stable over by Fact 2.3. It follows that is definable over (see e.g. [CG22, Proposition 3.15]). Since is the unique extension of definable over , it suffices to show that . Let and . Clearly and and so . Since is idempotent, we conclude that and in particular . For any we have
hence . ∎
2.5. Abelian groups
In this section we give a positive answer to Problem 2.10 in the case of abelian groups (in arbitrary theories):
Proposition 2.18.
Assume that is an abelian group and is a generically stable idempotent type. Then is generically transitive.
Proof.
Let be any small model such that is -invariant. Let be given, and assume towards contradiction that , then by Fact 2.1(4). Let be such that and forks over . We extend to a Morley sequence . For each , let . Then we have:
-
(1)
, hence by idempotence of we have , and so , and , so , in particular ;
-
(2)
for any permutation of , we have (by Fact 2.1(1));
-
(3)
in particular, for every we have
-
(4)
and (as is abelian);
-
(5)
hence for every .
Thus, by (1) and (5), for every we have , and by compactness the set is consistent. But this contradicts the choice of by Fact 2.1(5). Hence , and since was an arbitrary small model (over which is invariant), we conclude the proof. ∎
Remark 2.19.
It was pointed out to us by Martin Hils that this argument is related to [HRK19, Lemma 5.1], which is used there to find idempotent types in abelian groups of finite -weight.
Remark 2.20.
In the case of an arbitrary group, the proof of Proposition 2.18 gives the following:
-
•
If is invariant and idempotent, then for any we have .
Indeed, this time we assume towards a contradiction that . We extend to a Morley sequence in over . Let be arbitrary, and let . Then we get for all , which gives a contradiction as in the proof of Proposition 2.18. To see this, note that since is idempotent, , , and . Hence , and so .
Remark 2.21.
Note that in Remark 2.20 we only assumed that the idempotent type is invariant. However, the assumption of generic stability is necessary in Proposition 2.18. Indeed, let , , and let be the unique global definable (over ) type extending . Then is idempotent (see [CG23, Example 4.5(1)]). But if we have , hence , so .
2.6. Inp-minimal groups
A similar argument with (local) weight applies to arbitrary inp-minimal groups. Recall that a type-definable group is inp-minimal if (we refer to [Adl07] and [Che14, Section 2] for the definition and basic properties of burden). In particular, every dp-minimal group is inp-minimal. Note that there exist dp-minimal groups which are not virtually abelian [Sim03]. Answering [CPS14, Problem 5.9], it was recently proved in [Sto23] and [Wag24] that all dp-minimal groups are virtually nilpotent.
We will use the following fact:
Fact 2.22.
[GOU13, Theorem 2.9] Let be generically stable over a small set and for some cardinal . Let be such that for all . Then .
Proposition 2.23.
Assume that is an arbitrary type-definable group which is inp-minimal. If is idempotent and generically stable, then is generically transitive.
Proof.
Assume that is inp-minimal and is a generically stable idempotent type. Let be a monster model with respect to and (the unique extension of to a type in which is invariant over ). Then is idempotent and generically stable over .
Let . Let , by assumption . As , by Fact 2.22 we must have at least one of the following (by idempotence and Fact 2.1(4)):
-
(1)
, hence ;
-
(2)
or , hence .
In both cases we obtain that is generically transitive (see Remark 2.12). ∎
2.7. Stable theories
In this section we provide a proof that all idempotent types in stable groups are generically transitive. This was known from [New91] (see also e.g. [BMPW16, Lemme 1.2] and references there), and recently generalized from idempotent types to idempotent Keisler measures in stable theories in [CG22] (see the discussion in Section 3.10). We provide two detailed proofs since in the following sections we will extend them to the case when is a stable type in an arbitrary theory, and also to the case of a generically stable type in a simple or even rosy theory.
The proof uses local stratified ranks:
Definition 2.25.
Let be an -type-definable group and as in Section 2.2. To a formula (with an arbitrary tuple of variables) we associate a formula . For , put . Finally, let . We consider the usual notion of -rank denoted by and -multiplicity denoted by .
The proofs of all items except (3) in the following fact are standard arguments as for usual -ranks (see e.g. [Pil96, Chapter 1]). The proof of (3) uses the choice of and is left as an exercise.
Fact 2.26.
Assume is stable. Then we have:
-
(1)
;
-
(2)
;
-
(3)
is invariant under left translations by the elements of ;
-
(4)
For any and we have that does not fork over if and only if for every ;
-
(5)
for any complete type over a model .
Remark 2.27.
Note that (5) follows from (4) and stationary of types over models in stable theories, but for -saturated it can also be shown easily directly, without using forking.
In the proof below, the role of is played by .
Proposition 2.28.
If is stable and is idempotent, then is generically transitive.
Proof.
By stability, is generically stable over some small . Let in . By Remark 2.12, it suffices to show that , where
Method 1 (without using forking). Suppose for a contradiction that , witnessed by a formula . Since is a generically stable idempotent and , we get . On the other hand, since is invariant over and the ranks are invariant under automorphisms, a formula with and can be mapped by an automorphism over to a formula with and . By Fact 2.26(3), we also have . Summarizing, , and . Hence, , a contradiction.
Method 2 (using forking). Since does not fork over , by Fact 2.26(4), we get for every . By Fact 2.26(3), we also have . Thus, . On the other hand, since is a generically stable idempotent and , we have . We conclude, using Fact 2.26(4), that does not fork over , and so it is the unique non-forking extension of which is equal to . ∎
2.8. Stable types in arbitrary theories
Method 1 from the proof of Proposition 2.28 extends to the case when is stable over , that is is -invariant and is stable (in a not necessarily stable theory). Recall that is stable if there are no sequences in and in such that for all and for some we have for all . It follows from the definitions that if is stable over , then it is generically stable over .
We refer to e.g. [ACP14, Section 1] and [HO10, Section 2] for the basic properties of stable types. Method 1 applies when is stable over because in this case , and we also have item (3) of Fact 2.26 (without any assumptions on ). Thus, Method 1 yields:
Proposition 2.29.
If is idempotent and stable over some small , then is generically transitive.
In order to prove this proposition using Method 2 from the proof of Proposition 2.28, we have to be careful with item (4) of Fact 2.26. Modifying a standard proof of item (4) (see e.g. [Pil96, Lemma 3.4]) gives the following weaker variant:
Fact 2.30.
Assume and is stable for a small model . Let be a set of realizations of and . Then for any extending we have that does not fork over if and only if for every .
While adapting the proof of [Pil96, Lemma 3.4] for Fact 2.30, the only essential difficulty is to show that if a formula does not fork over (where ), then some positive Boolean combination of -conjugates of is definable over . For that one needs to use the assumption that is a set of realizations of in order to have that is stable, which allows to use symmetry of forking and -definability of non-forking extensions of .
The following is a strengthening of the fact saying that is the unique non-forking extension of to a global type, and follows by one of the standard proofs:
Proposition 2.31.
Assume that is generically stable, and (where is a monster model with respect to ). Let be an extension of such that does not fork over for every small set of independent realizations of . Then .
More generally, in the assumption, it is enough to consider only the sets of independent realizations of containing a fixed realization of in .
Proof.
We will deduce the proposition from the following claim.
Claim.
If for all , then is a Morley sequence in . In the more general version of the assumption, the same holds but assuming that .
Proof.
This is induction on . The base step is trivial, as by assumption.
Induction step. Consider any in for all . By induction hypothesis for all . The goal is to prove that , equivalently .
Suppose for a contradiction that but . Extend to a Morley sequence in by . Since is a Morley sequence in , we get, by generic stability of , that the formula divides over for every . As by assumption we know that does not fork over , we conclude that for all .
Pick any . By the above conclusion, for all , so, by generic stability of , .
On the other hand, since and is Morley sequence in over , we get that for all . Since is a Morley sequence in , and as such it is totally indiscernible, we get that for all . Therefore, .
Summarizing, the last two paragraphs yield . This contradicts the -invariance of , because as both these elements satisfy . ∎
Now consider any . Pick in for all , and put . By the claim, is a Morley sequence in . Since for all , by generic stability of , we get . Thus, . ∎
Now, to prove Proposition 2.29 via Method 2, using Fact 2.30, we get that for every set of realizations of (where ). This implies (by Fact 2.26(3) — which does not require any assumptions) that for every set of realizations of containing . Since (by the idempotence of ), we conclude, using Fact 2.30, that does not fork over for every set of realizations of containing . Hence, by Proposition 2.31.
2.9. Simple theories
Assume that is a simple theory, and as usual that is idempotent and generically stable over . Method 2 extends from stable to simple theories using stratified Shelah degrees in place of stratified local ranks. In simple theories, they are wrongly defined in Definition 4.1.4 of [Wag00] (as they are not left-invariant due to the lack of associativity outside ). A way to fix it is to use suitable (as in Definition 2.25) or to apply Definition 4.3.5 of [Wag00] (in the special case of type-definable rather than hyper-definable groups). In any case, by [Wag00], stratified Shelah degrees satisfy items (1)–(4) of Fact 2.26, so Method 2 applies directly and yields the following generalization of Proposition 2.28:
Proposition 2.32.
Let be simple, an -type-definable group, and idempotent and generically stable. Then is generically transitive.
2.10. Rosy theories
In the case of groups in rosy theories, again we can apply Method 2, using stratified local thorn ranks. They were defined and studied in [EKP08] (see [EKP08, Definition 1.13]) in the case of definable groups, and extend easily to type-definable groups (using as in Definition 2.25). By [EKP08], stratified local thorn ranks satisfy items (1)–(4) of Fact 2.26, with thorn forking in place forking in item (4). However, [GOU13, Theorem 3.4] tells us that if a type is generically stable and , then forks over if and only if thorn forks over . We have all the tools to prove the following generalization of Proposition 2.32 via Method 2.
Before its statement, let us first recall local thorn ranks and define stratified local thorn ranks :
Definition 2.33.
For a finite set of partitioned formulas with object variables and parameter variables , a finite set of formulas in variables , and natural number , the -rank is the unique function from the collection of all consistent formulas with parameters to satisfying: if and only if there is , some and parameter such that:
-
(1)
for infinitely many , and
-
(2)
is -inconsistent.
Given a (partial) type closed under conjunction we define to be the minimum of for .
Definition 2.34.
-
(1)
For a formula , let , where is chosen in Section 2.2.
-
(2)
For a finite set of formulas in variables , put . For a finite set of formulas in variables , put .
-
(3)
The stratified -rank is defined as the unique function satisfying:
if and only if there is a formula , some and parameters and anywhere such that:-
(a)
for infinitely many , and
-
(b)
is -inconsistent.
Given a (partial) type closed under conjunction we define to be the minimum of for .
-
(a)
Proposition 2.35.
If is rosy and is a type-definable group, then every generically stable idempotent type is generically transitive.
Proof.
Assume is generically stable over and let in and . As usual, by Remark 2.12 it suffices to show that . Since does not fork over , it does not -fork over , so for every finite , , . Since , we conclude that . On the other hand, since is a generically stable idempotent and , we have . Therefore, does not -fork over . In order to conclude that does not fork over (and so coincides with , which finishes the proof), by virtue of [GOU13, Theorem 3.4], it remains to show the following:
Claim.
is generically stable over .
The claim follows from Fact 2.3 since is generically stable over (and so over ) and is realized by for . ∎
2.11. Generic transitivity of and stratified rank localized on
As we saw in Sections 2.7–2.10, generic transitivity of an idempotent generically stable type can be established using a well-behaved stratified rank. In this section, working in an arbitrary theory, we define an analog of the stratified rank in stable theories (Definition 2.25) restricting to formulas with parameters from a Morley sequence in a generically stable type (aiming for it to satisfy the properties similar to Fact 2.26 needed to apply Method 1 or 2 from the proof of Proposition 2.28). For technical reasons, we will make a stronger assumption that is generically stable for all . In NIP theories, or even in NTP2 theories (by [CGH23]), this follows from generic stability of ; but it is open in general if generic stability of implies that is generically stable (a counterexample was suggested in [ACP14, Example 1.7], however it does not work — see [CGH23a, Section 8.1]). As the main result of this section, we will show that this rank is left invariant (under multiplication by elements from ) if and only if is generically transitive.
The following proposition will be used to show that our ranks are finite, but it may be of independent interest.
Proposition 2.36.
Let , a small model, and assume that is generically stable over for every . Let be finite and let be a Morley sequence in over . Let be any formula (possibly with parameters, and arbitrary tuples of variables). Then there are only countably many types in .
Proof.
As there are only finitely many possibilities for substitutions of the elements of the finite set in place of some variables in , without loss of generality we may assume that .
Denote , say with corresponding to the free variables of for . The next claim follows from generic stability of .
Claim.
There exists such that: for every there is a subset of cardinality and such that for every pairwise distinct we have .
Proof.
By generic stability of , there is such that for every Morley sequence in over and for every , either for all but many ’s we have , or for all but many ’s we have . Put , suppose that it does not satisfy the requirement in the claim, and choose a witness for that. Then, by recursion on , we can find pairwise distinct numbers and for and so that and for all . As
is a Morley sequence in over , the previous sentence contradicts the choice of . ∎
By the claim, varying , we have only countably many possibilities for the finite set and two possibilities for , so only countably many possibilities for .
For , let . The proof of the proportion is by induction on .
Base step (). For any , the type is determined by the pair and the truth values of finitely many sentences for . So we get countably many possibilities for .
Induction step (). By induction hypothesis , so
for any and . Since is determined by the pair together with
we conclude that there are countably many types in . ∎
Now let be an -type-definable group and such that is generically stable over for all . Let be a bigger monster model. We define a version of stratified local ranks, where the inconsistent types witnessing the increase in rank have to be defined over a Morley sequence in .
Definition 2.37.
-
(1)
For a formula , let , where is chosen in Section 2.2.
-
(2)
Following Shelah’s terminology (rather than Pillay’s from [Pil96]), given , by a -formula over we mean a formula of the form or for from .
-
(3)
By a -type over we mean a consistent collection of -formulas over . Two such types are explicitly contradictory if there is a -formula contained in one of these types such that the negation of this formula is in the other type.
Definition 2.38.
-
(1)
We define a function from the collection of all partial types in over to as a unique function satisfying: if and only if for every finite and there exist pairwise explicitly contradictory -types whose parameters altogether form a Morely sequence in over together with the parameters of and such that for all .
-
(2)
If , then is defined as the maximal number such that for every finite with there are pairwise explicitly contradictory -types whose parameters form a Morley sequence in over together with the parameters of and such that for all .
Lemma 2.39.
Assume that is generically stable for all . Then the ranks have the following properties.
-
(1)
.
-
(2)
.
-
(3)
is invariant under automorphisms of fixing pointwise.
-
(4)
In the definition of , one can use in place of the finite piece .
-
(5)
.
Proof.
(1) If not, then, by compactness, there is a tree , where each is a -formula, branches are consistent, the sons of every node are pairwise explicitly contradictory, and the parameters of the types form a Morley sequence in over . So the number of complete -types over these parameters is which contradicts Proposition 2.36.
(2) and (3) are straightforward from the definitions.
(4) follows from (1) (namely, it is enough to consider only finite values of the rank), and compactness. One just gets that if and only if there is a tree of -types with suitable parameters and such that each branch is consistent with and the sons of every node are pairwise explicitly contradictory.
(5) follows from the observation that each but at most one of the types in the definition of implies . ∎
We consider the left action of on partial types over as follows:
Definition 2.40.
For and a partial type over a set , we define to be a partial type over defining the set .
Proposition 2.41.
Generic transitivity of is equivalent to the invariance of the ranks under action on the left by elements of .
Proof.
By Lemma 2.39(5), it suffices to show that for any and any partial type which implies we have .
By generic transitivity of and Remark 2.12 we have , where . Then the type-definable group is generically stable witnessed by , so by Fact 2.5. Hence, . Thus, since is equivalent to (by the choice of ), it is enough to show that . For that, by induction on , we will show that implies that .
The base step is obvious. For the induction step, consider any and pairwise explicitly contradictory -types whose parameters form a Morley sequence in over together with the parameters of and such that for all . Without loss of generality we may assume that this Morley sequence is over . By induction hypothesis, for all . On the other hand, by the choice of and , for any and any in we have is equivalent to and is equivalent to . Thus, for all for some pairwise explicitly contradictory -types whose parameters form a sequence , where for every , or . It remains to show that is a Morley sequence in over . So we need to show that .
If , then this follows from the fact that . Consider the case . Since , we have and so by left transitivity. On the other hand, by generic transitivity of and Remark 2.12, . Hence, (by Fact 2.1(2)).
Let in , and suppose for a contradiction that . By Proposition 2.31, there is a formula with being a Morley sequence in (over ) such that but . Then we follow the lines of Method 1, using in place of and left invariance of under (which follows by assumption) together with invariance under . At the end, we get , and , which contradicts the choice of . ∎
Problem 2.42.
Does the equivalence in Proposition 2.41 hold only assuming that is generically stable? (As opposed to is generically stable for all .)
2.12. Stabilizer of a generically transitive type is an intersection of definable groups
As before, let be a type-definable group and let be a generically stable type over . By Proposition 2.7, we know that is an intersection of relatively -definable subgroups of .
Question 2.43.
Assuming that is generically stable and idempotent, is an intersection of -definable groups?
This question is open only in the case of type-definable (in the definable case, the answer is trivially positive by the above comment).
Proposition 2.44.
If is type-definable and is generically stable, idempotent and generically transitive, then the answer to Question 2.43 is positive.
Proof.
This proof is an elaboration on the proof of Hrushovski’s theorem that a type-definable group in a stable theory is an intersection of definable groups [Hru90].
Choose as in Section 2.2, we will only use that and is defined and associative on and for all .
We will prove that there exists an -definable set such that and is a group. Then and , where ranges over the formulas chosen as above. Therefore, by Proposition 2.7 and compactness, is an intersection of -definable groups (all with group operation given by ), so the proof will be finished.
We can clearly assume that the partial type is closed under conjunction. For any put . Let be the -definition of ; this is a formula over , since is definable over . Also, implies .
Claim 1.
.
Proof.
Consider any , i.e. . Take . Then . Hence, for any we have , and so .
Consider any . Take . Then for every we have , so . Hence, , so (the equality holds, as and ), whence . ∎
By Claim 1, choose such that for all .
Claim 2.
For every and we have .
Proof.
Consider any and as above. Take any realizing . By generic transitivity and Remark 2.12, , and so . On the other hand, by left transitivity, . Hence, by transitivity of forking for generically stable types (Fact 2.1), and so . Therefore, since , we get . Since , this means that . On the other hand, . Moreover, since and (the last inclusion holds by Claim 1), we have . So we conclude that . ∎
Put and . By Claim 2, , and clearly . Both and are -definable. Using the choice of and the inclusions , one easily checks that is closed under . Finally, put . It is clearly -definable.
By generic transitivity of (and Remark 2.11), the type-definable group is generically stable witnessed by (Definition 2.4), so by Fact 2.5. Therefore, as we have seen above that , we get that . Summarizing, .
One easily checks, using the choice of and the inclusion , that is closed under . Since , we have that is associative on and is neutral in . Also, for any there is with which implies that . Hence, is a group. ∎
Remark 2.45.
If we drop the assumption that is generically transitive, then the main difficulty in adaptation of the above proof is in Claim 2. Namely, by exactly the same method we only get “For every and we have ”.
2.13. Chain conditions for type-definable groups on a generically stable type
One approach towards Problem 2.10 is to adapt Newelski’s “2-step generation” theorem [New91, Theorem 2.3] that provides a positive answer for types in stable theories. One ingredient is the existence of a smallest type-definable group containing a given type in a stable theory. In this section we investigate this question (and corresponding chain conditions and connected components) for groups type-definable using parameters from a Morley sequence of a generically stable type.
The following is a generalization of [DK12, Lemma 2.1] to type-definable groups (combining it with [HP18, Lemma 2.1]).
Lemma 2.46.
Let be an -type-definable group and a global type generically stable over a small set .
-
(1)
Assume that is a countable partial type over , and let . Then for any linear order , any tuple and any , if for all , then for any with we have
-
(2)
Assume that is generically stable over for all . Let with be a countable partial type over . Then satisfies the following: for any linear order , any tuple and any , if , where , for all , then for any with we have
-
(3)
Assume that is generically stable over for all . Let be a family of subgroups of that is invariant under -automorphisms and closed under (possibly infinite) intersections and supergroups (so for example could take to be all subgroups of ).
Then for any linear order , any and any , the intersection of all subgroups of from type-definable with parameters from is a subgroup of in relatively type-definable over .
Proof.
(1) Without loss of generality we may clearly assume that ; and extending to a longer Morley sequence if necessary, we may assume . First we show that there exists some with . Assume not, then by induction on we can choose and pairwise distinct so that
Let with contain the parameters of , and let be a countable sublanguage of containing all of the formulas in . By the choice of , applying Erdős-Rado and passing to a subsequence of we may assume that the sequence is “-indiscernible” with respect to , i.e. tuples have the same -type over for all , and either for all , or for all . Assume we are in the former case (the latter case is similar).
First assume that for some in , then for some . Hence, by -indiscernibility, we have for all . Taking sufficiently large, this contradicts generic stability of .
Otherwise we have , for all . Note that for any , we have (as ). By compactness there is a formula (without loss of generality the partial type is closed under conjunctions) so that: for any , . By -indiscernibility we then have: for any , for any , . Let now be any finite subset of , and let . As , it follows that for all , . Taking sufficiently large, this again contradicts generic stability of (in fact, even “generic NIP” of ).
We have thus shown that there exists with . Given an arbitrary , there exists a permutation of sending to . As is totally indiscernible over by generic stability of , there exists with for all . Applying we thus have:
(2) Let be arbitrary. Fix any . Let , so that and . As , we can choose tuples for so that: , for , , and for all . As is totally indiscernible over , it follows that , hence applying (1) to the family , , and generically stable type we conclude that
(3) Let be type-defined by a partial type with parameters in for some with a small linear order. We can write for some ordinal , with each a countable partial type relatively defining a subgroup of , and a countable subsequence of . In particular , as is closed under supergroups. Fix any . For any , applying (2) we have . And by indiscernibility of over , -invariance of and closure under intersections, we have . Hence . ∎
Question 2.47.
Does Lemma 2.46(2) hold only assuming that is generically stable? The answer is positive for the analog for definable groups instead of type-definable groups ([DK12, Lemma 2.1]), but the proof there does not immediately seem to generalize to the type-definable case. We also expect that an analog of Lemma 2.46 holds for invariant instead of type-definable subgroups, but do not pursue it here.
Corollary 2.48.
Let be an -type-definable group and a global type so that is generically stable for all . Assume that is invariant over a small set , and consider the family
Then the group is type-definable over , and .
Proof.
Let be as in Lemma 2.46, and fix some . Assume we are given an arbitrary small linear order and arbitrary , and let with be type-definable over . We can choose some . Let . Then the family is -invariant by -invariance of , and is closed under supergroups and intersections. Applying Lemma 2.46(3) with and the sequence , we find some type-definable over with . Applying Lemma 2.46(3) again to the sequence , we find some type-definable over with . This shows that the group is type-definable, over . But it is also -invariant (since the family is -invariant by -invariance of ). Hence is type-definable over , and . ∎
Question 2.49.
-
(1)
Does depend on ?
-
(2)
Is an intersection of definable groups?
-
(3)
Let be a generically stable type (idempotent and such that is generically stable for all , if it helps). Does there exist a smallest type-definable (over a small set of parameters anywhere in ) group with ?
All of these questions have a positive answer assuming is idempotent and generically transitive:
Proposition 2.50.
Assume that is generically stable, idempotent and generically transitive. Then is the smallest type-definable (with parameters anywhere in ) subgroup of so that .
Moreover, if is generically stable for all , then for any small so that is invariant over , and is an intersection of definable groups.
Proof.
We note however that we cannot answer Question 2.49(3) positively by proving a chain condition for groups containing and (type-)definable with arbitrary parameters:
Example 2.51.
Let be a monster model of the theory ACVF of algebraically closed fields, and let be the additive group. For and , let be the closed (respectively, open) ball of radius with center . Fix , given a ball , by quantifier elimination in ACVF, density of , and the fact that for any two balls either one is contained in the other, or they are disjoint,
determines a complete type , called the generic type of .
Now let and . As for any we have , hence for any we have and if , then for any ball . And if then , so we have , and . In particular , and is left--invariant. Note that generically stable over any with (as it is internal to the residue field, or can see directly that it commutes with itself). Hence is idempotent.
Note that is a large strictly descending chain of definable subgroups of with for all . So we have arbitrary long descending chains of definable subgroups of containing (but the smallest one is still definable).
3. Idempotent generically stable measures
3.1. Overview
The main aim of this section is to prove the following:
Theorem.
Remark 3.1.
In particular, if is NIP and is abelian, there is a one-to-one correspondence between generically stable idempotent measures and type-definable fsg subgroups of .
Remark 3.2.
The assumption that is fim cannot be relaxed.
Indeed, consider as a definable group in , and let be a monster model. Let be the normalized Haar measure on and the type of the cut above . Let be the unique smooth extension of to the monster model, the unique definable extension of , and . Then is definable and idempotent. But , in particular , so cannot satisfy the conclusion of the theorem by Remark 3.39 (see also [CG23, Example 4.5]).
This section of the paper is organized as follows. We briefly recall the setting and some properties of Keisler measures in Section 3.2. In Section 3.3 we recall some basic facts and make some new observations involving fim measures — they provide a generalization of generically stable measures from NIP to arbitrary theories in the same way as generically stable types in the sense of [PT11] provide a generalization from NIP to arbitrary theories. In Section 3.4 we prove that the usual characterization of generic stability — any Morley sequence of a fim measure determines the measure of arbitrary formulas by averaging along it — holds even when the parameters of these formulas are themselves replaced by a measure, see Theorem 3.13. In Section 3.5 we collect some basic facts about definable pushforwards of Keisler measures. In Section 3.6 we develop some theory of fim groups (i.e. groups admitting a translation invariant fim measure) in arbitrary theories, simultaneously generalizing from fsg groups in NIP theories and generically stable groups in arbitrary theories (in particular, that this translation invariant measure is unique and bi-invariant, Proposition 3.32). In Section 3.7 we develop an appropriate analog of generic transitivity for fim measures, generalizing some of the results for generically stable types from Section 2.4. Finally, in Section 3.8, we put all of this together in order to prove the main theorem of the section, adapting the weight argument from Section 2.5 to a purely measure theoretic context.
In Section 3.9 we isolate a weaker property of support transitivity and connect it to the algebraic properties of the semigroup induced by on the support of an idempotent measure. In Section 3.10 we illustrate how the Keisler randomization can be used to reduce generic transitivity to support transitivity in stable groups, and discuss more general situations when the randomization may have an appropriate stratified rank.
3.2. Setting and notation
We work in the same setting as in Section 2.2. For a partitioned formula , is the same formula but with the roles of the variables swapped. In the group setting, if , then . Let . Then a Keisler measure (in variable over ) is a finitely additive probability measure on (modulo logical equivalence). We denote the collection of Keisler measures (in variable over ) as . Given , we let denote the support of , i.e. the (closed) set of all such that for every . We refer the reader to [CG22, CG23] for basic definitions involving Keisler measures (e.g. Borel-definable, definable, support of a measure, etc). Given a partial type over , we will consider the closed set of measures supported on .
3.3. Fim measures
Throughout this section we work in an arbitrary theory , unless explicitly specified otherwise.
Definition 3.3.
Let , and suppose that is Borel-definable. Then we define the Morley product of and , denoted , as the unique measure in such that for any , we have
where:
-
(1)
is -invariant and contains all the parameters from ,
-
(2)
is defined by for some (equivalently, any) in ,
-
(3)
is the unique regular Borel probability measure on corresponding to the Keisler measure .
See e.g. [CG22, Section 3.1] for an explanation why this product is well-defined and its basic properties. We will often abuse the notation slightly and replace with either or simply when it is clear from the context, and sometimes write as .
Definition 3.4.
Let and suppose that is Borel-definable. Then we define , , and . (In general, need not be commutative/associative on Borel definable measures in arbitrary theories.)
Definition 3.5.
[HPS13] Let and a small model. A Borel-definable measure is fim (a frequency interpretation measure) over if is -invariant and for any -formula there exists a sequence of formulas in such that:
-
(1)
for any , there exists some satisfying: for any , if then
-
(2)
.
We say that is fim if is fim over some small .
Remark 3.6.
In NIP theories, fim is equivalent to each of the following two properties for measures: dfs (definable and finitely satisfiable) and fam (finitely approximable [CS21]), recovering the usual notion of generic stability for Keisler measures [HPS13]. Outside of the NIP context, fim (properly) implies fam over a model, which in turn (properly) implies dfs (see [CG20, CGH23a]).
Generalizing from generically stable measures in NIP, one has:
Fact 3.7.
[CGH23a] If is fim and is Borel definable, then .
Definition 3.8.
[CGH23] Let and a small submodel such that is -invariant, and . We say that is self-averaging over if for any measure with and any formula we have
The following generalizes a standard characterization of generically stable measures in NIP theories [HPS13] to fim measures in arbitrary theories (and demonstrates in particular that if is fim viewed as a Keisler measure, then it is generically stable in the sense of Section 2.1; indeed, is generically stable over if for every Morley sequence in over we have — and if is self-averaging, this property holds, see [CG20, Proposition 3.2]):
Fact 3.9.
[CGH23, Theorem 2.7] If and is fim over , then is self-averaging over .
The following would be a natural generalization of stationarity for generically stable types to measures:
Conjecture 3.10.
Let be fim over and let be a small set with . If is -invariant, Borel-definable, and , then .
Conjecture 3.10 is known to hold when is a type and is an arbitrary theory (by Fact 2.1(2)) and when is a measure but is NIP (by [HPS13, Proposition 3.3]). The following proposition is a special case of Conjecture 3.10 sufficient for our purposes here.
Proposition 3.11.
Let be fim over and . Suppose is -invariant and . If either of the following holds:
-
(1)
is definable,
-
(2)
or the measures are fim for each and is Borel-definable;
then .
Proof.
The only difference between the two cases is the justification of equation below. Suppose that we are given some with the described properties. Since is -invariant and (Borel-)definable, it follows that is (Borel-)definable over .
Claim.
We have that (in either of the two cases).
Proof.
By assumption . Assume that we have already established for some . Fix an arbitrary formula and . Let . Since the measure is definable over , there exist formulas , and such that
() |
Then we have:
with the following justifications for the corresponding steps:
-
(a)
induction hypothesis;
-
(b)
in Case (1), is fam over since is fam over and (and fim implies fam over a model), by [CG20, Proposition 2.10(b)], is definable over , and fam measures commute with definable measures [CGH23a, Proposition 5.17]; in Case (2), is fim over , is Borel-definable over , and fim measures commute with Borel-definable measures [CGH23a, Proposition 5.15];
-
(c)
by ;
-
(d)
by assumption;
-
(e)
is fim, and fim measures commute with all Borel-definable measures [CGH23a, Proposition 5.15] (alternatively, fam measures commute with definable measures).
As and were arbitrary, we conclude . And then . ∎
Let now be arbitrary. Let , by the claim above we have . Note that for every , . As is fim over , it is self-averaging over by Fact 3.9, hence . ∎
3.4. Fim measures over “random” parameters
In this section we prove a generalization of Fact 3.9 of independent interest, demonstrating that any Morley sequence of a fim measure determines the measure of arbitrary formulas by averaging along it — even when the parameters of these formulas are allowed to be “random”. More precisely:
Theorem 3.13.
Let be fim over , , , , and . Suppose that is arbitrary such that and . Then
Moreover for every there exists so that for any as above, we have for all but many .
Remark 3.14.
Fact 3.9 corresponds to the special case when is a type. We note that this result is central to the proof of our main theorem, and is new even for NIP theories.
Our proof of Theorem 3.13 relies on the use of Keisler randomization in continuous logic, as introduced and studied in [BYK09]. We will follow the notation from [CGH23, Section 3.2]. Let be a complete first order theory. We let denote the (continuous) first order theory of its Keisler randomization (we refer to [BYK09, Section 2] for the details). Let be a model of and let be a probability algebra. We consider the model of , which we usually denote as for brevity, defined as follows. We let
equipped with the pseudo-metric . Then is constructed by taking the metric completion of and then identifying random variables up to -measure . We let be the set of classes of elements of . By construction, is a metrically dense (pre-)substructure of .
Let be a monster model of such that . The model is almost never saturated, so we will always think of as (elementarily) embedded into a monster model of , i.e. . If , we let denote the constant random variable taking value , i.e. is the equivalence class of the maps which send to the point (equivalence up to measure ). If , we let . If is an -formula, we let denote the corresponding continuous formula in the randomization. This formula is evaluated on tuples of elements from via
and is extended to via uniform limits. For , will denote the space of types in the tuple of variables over in .
Remark 3.15.
Note that for any , there exists a finite -measurable partition of with the property that for each the function is constant on each element of . Given such an and , we write for the tuple of constant values of the functions in on the set . Note that for each , is an element of .
The following fact can be derived from basic facts about continuous logic:
Fact 3.16.
Suppose that . Then there exists a net of tuples where such that .
The following observations were made by Ben Yaacov in an unpublished note [BY09]. For a detailed verification, we refer the reader to [CGH23, Section 3.2].
Fact 3.17.
Let be a monster model of , , and .
-
(1)
There exists a unique type such that for any -formula , , and any measurable partition such that each element of is constant on each element of ,
-
(2)
If is definable, then there exists a unique type such that:
-
(a)
;
-
(b)
is definable over ; if is -definable, then is -definable.
-
(a)
Remark 3.18.
The claims in Fact 3.17 hold in the context where is an infinite tuple of variables. The infinitary results follow easily from their finite counterparts.
Corollary 3.19.
Suppose that , , and . Then
Proof.
Fix and . Fix a finite measurable partition of such that each element of is constant on each element of . From the definitions we have:
By quantifier elimination in , we conclude that . ∎
We recall some results from [CGH23] connecting the randomized measures, the Morley product and generic stability in and . These are [CGH23, Proposition 3.15], [CGH23, Corollary 3.16] and [CGH23, Corollary 3.19], respectively.
Fact 3.20.
Corollary 3.21.
If is a definable measure then .
Proof.
First note
We want to show that . By quantifier elimination in , it suffices to show that for every -formula and , we have that . By Fact 3.16, fix a net of elements each in such that . For each , choose a finite measurable partition of such that each element of is constant on each element of . We have the following computation (using Fact 3.17):
where the first and last equality follow from the fact that and are continuous maps, by definability of , hence of and , and Fact 3.17(2)(b). ∎
The following fact is [CGH23, Lemma 3.13].
Fact 3.22.
Suppose that and is a net of elements such that and . For each , let be a finite measurable partition of such that each element of is constant on each . Then
where the limit is calculated in the space .
Proposition 3.23.
Let be fim over a small model , and with of the same sort as for all . Suppose that such that and . Then for any we have
Proof.
The measures , and can be associated to complete types , and , respectively, by Fact 3.17(1). As is -definable, is also -definable, and so the types and are well-defined and definable over by Fact 3.17(2). We then have:
-
(1)
is generically stable over (by Fact 3.20(3));
-
(2)
(by Corollary 3.21);
-
(3)
and (by Corollary 3.19).
Let and in be so that . Then is a Morley sequence in over by (2) and (3) from above. By Fact 3.16, choose a net of tuples in such that . For each , choose a finite measurable partition of such that each function in is constant on each element of . Now, given any , we then have the following computation:
where the corresponding equalities hold for the following reasons:
-
since the type is generically stable over by (1) and is a Morley sequence in over ;
-
by the choice of and, as is definable over by (1), the map is continuous;
-
by the definition of (Fact 3.17);
-
for a fixed , the computations of the left hand side and the right hand side are the same;
-
since is definable, the map is continuous, hence the map is continuous;
-
by Fact 3.22. ∎
We will use the following general topological fact [CGH23, Lemma 2.3]:
Fact 3.24.
Let be an arbitrary function from a compact Hausdorff space to a compact interval . Suppose there is a closed subset satisfying the following properties:
-
(1)
the projection of onto is all of ;
-
(2)
if and is strictly increasing, then ;
-
(3)
for any , .
Then is continuous and, for any , there is an such that: for any , .
Finally, we can derive the main theorem of the section:
Proof of Theorem 3.13.
The “moreover” clause of Theorem 3.13 for formulas without parameters follows from Proposition 3.23 by compactness (using Fact 3.24).
Namely, first let be arbitrary. We let
The assumptions of Fact 3.24 are satisfied. Indeed, (1) holds since for every , gives an element in projecting onto it. (2) For every strictly increasing we have a continuous map defined by . Now if then still , where is the pushforward of the measure by (see Definition 3.25). And (3) holds by Proposition 3.23. Then we obtain the required applying Fact 3.24.
Now assume we are given , , and is an arbitrary formula with parameters, say of the form for some and . Let be as given by the above for the formula without parameters. Given any with and , consider the measures defined by , and defined by . Note that . By the choice of and the previous paragraph (with and in place of and ) we have
3.5. Definable pushforwards of Keisler measures
We record some basic facts about definable pushforwards of Keisler measures.
Definition 3.25.
Let be a definable map. For , we define the push-forward measure in , where for any formula .
Proposition 3.26.
Let . Let be a small set and let be an -definable map. Then we have the following:
-
(1)
if is -invariant, then is -invariant;
-
(2)
if is -definable, then is -definable;
-
(3)
if is fim over , then is fim over ;
-
(4)
, and if is a bijection then these sets are equal.
Proof.
(1) Straightforward.
(2) Note that for any formula , we have where .
(3) Let be fim over .
Claim.
For any ,
Proof.
We prove the claim by induction on . The base case is trivial. Assume the claim holds for . For , let be the -definable map given by . Note that induces a pushforward from to . Then our induction hypothesis says . Fix and let
Let be a small model containing and all relevant parameters. Then
We now show that is fim over . Fix a formula in . Since is fim, let be a sequence of -formulas witnessing this for the formula as in Definition 3.5. To avoid “scope-of-quantifiers” confusion, we let be new variables with of the same sort as for each . For each , we consider the -formula
Note that
() |
We will show that the formulas witness that is fim over with respect to the formula .
Fix . Then there exists some so that: for any and any with we have
() |
Now, suppose that is such that . Then, by definition of , there exists some such that and . Therefore,
Finally, by the Claim and we have
We conclude is fim.
(4) Immediate from the definitions. ∎
3.6. Fim and fsg groups
Throughout this section, will be a -type-definable group. Let be chosen for as in Section 2.2.
Definition 3.27.
[HPP08, Definition 4.1] A (-)type-definable group is fsg (finitely satisfiable generics) if there is some and small such that for every , is finitely satisfiable in .
We consider a natural generalization of generically stable groups (Section 2.3) to fim groups.
Definition 3.28.
-
(1)
We let denote the (closed) set of all measures supported on , i.e. with .
- (2)
Definition 3.29.
We will say that a (-)type-definable group is fim if there exists a right -invariant fim measure , i.e. for all .
Remark 3.30.
-
(1)
In any theory, if is fim then it is both definably amenable and fsg.
-
(2)
If is NIP and is fsg, then it is fim.
Proof.
(1) Indeed, let witness that is fim. In particular, is finitely satisfiable in some small (Remark 3.6), and for all . Take any . Then , and also , hence finitely satisfiable in , for all .
Problem 3.31.
Does fsg imply definable amenability without assuming NIP? Do there exist fsg (and definably amenable) groups that are not fim?
The following is a simultaneous generalization of Fact 2.5 from types to measures in arbitrary theories, and of the previously known case for measures under the NIP assumption [HPS13, Theorem 4.3].
Proposition 3.32.
Suppose that is a -type-definable fim group, witnessed by a right--invariant fim measure . Then we have:
-
(1)
;
-
(2)
is left -invariant;
-
(3)
is the unique left -invariant measure in ;
-
(4)
is the unique right -invariant measure in .
Proof.
(1) Fix a formula . Let be a small model such that is -invariant and contains the parameters of . As is fim over , is also fim over by Proposition 3.26(3).
Suppose (so ). Then, for (in the notation of Section 3.2) we have (for any ):
where follows by right -invariance of applied to the formula .
Using this and that fim measures commute with all Borel definable measures (Fact 3.7) we have:
Similarly, for and any (so ), by right -invariance of we have . Then
We conclude that .
(2) For any and , using (1) and right -invariance of for the formula we have:
(3) Suppose that is left -invariant, and let be arbitrary. Let be a small model such that is invariant over and contains the parameters of .
As in (1), for any (so ), by right -invariance of we have . Hence
(*) |
Second, for any (so ), by left -invariance of we have (for any ):
(†) |
So the map is constant on the support of , hence Borel. As is fim, the proof of [CGH23a, Proposition 5.15] implies
(‡) |
where on the left we view as a regular Borel probability measure restricted to the compact set . Then
(4) Let be right -invariant, and containing the parameters of and such that is -invariant.
As in (3), the map is -measurable since it is constant on the support by right -invariance of . As is fim we can apply [CGH23a, Proposition 5.15] again to get
As is left -invariant by (2), for any , we have (for ):
Combining we get
Proposition 3.32 and its symmetric version with “left” swapped with “right” (obtained by a “symmetric” proof) yield:
Corollary 3.33.
An ()-type-definable group is fim if and only if there exists a left -invariant fim measure .
3.7. Idempotent fim measures and generic transitivity
Let be a -type-definable group, and we are in the same setting and notation as in Section 2.2. For a measure , we let
denote the right-stabilizer of .
Fact 3.34.
When is a measure definable over , then is an -type-definable subgroup of (see e.g. [CG22, Proposition 5.3]).
Definition 3.35.
Suppose that is Borel-definable. Then for any measure , the (definable) convolution of and , denoted , is the unique measure in such that for any formula ,
We say that is idempotent if .
When is NIP, it suffices to assume that is invariant (under for some small model ), as then is automatically Borel-definable ([Sim15]). We refer to [CG22, Section 3] for a detailed discussion of when convolution is well-defined.
We now consider the main question investigated in [CG22, CG23] in the case of measures (generalizing from types in Section 2.4).
Again, we let .
Definition 3.36.
For the rest of the section, we let be the (-definable) map (where is viewed as a globally defined function whose restriction to defines the group operation, see Section 2.2).
Proposition 3.37.
Let be an idempotent fim measure. Then the following are equivalent:
-
(1)
;
-
(2)
;
-
(3)
for every .
Proof.
(2) (3) This equivalence only uses that is a definable measure.
By the definition of we have for all , where . By definition of we have: if and only if there exists a formula and a small model so that contains the parameters of and is -definable, such that
where is the restriction of to , is the restriction viewed as a regular Borel measure on . By definability of , the function is continuous (and non-negative), hence the integral is if and only if for some , that is for some and , that is for some .
(3) (1) Fix , let be such that is -invariant, and let .
Claim 3.
We have .
Proof.
Any is of the form for some . By (3) we have:
Therefore,
∎
Claim 4.
The measure is -definable.
Proof.
Since is -definable, it is also -definable. Consider the -definable map . Note that and so by Proposition 3.26, is also -definable. ∎
Hence, by Proposition 3.11 over , we get . That is, , so .
(1) (3) Let be arbitrary, and contain its parameters such that is -invariant. The measure is right -invariant, and given any we have by (1). So for we have , hence . ∎
Definition 3.38.
We say that an idempotent fim measure is generically transitive if it satisfies any of the equivalent conditions in Proposition 3.37.
In particular, a type is generically transitive in the sense of Definition 2.13 if and only if, viewed as a Keisler measure, it is generically transitive.
The following is a generalization of Remark 2.11:
Remark 3.39.
Proof.
(1) is a fim group, witnessed by the right -invariant fim measure .
(2) is type-definable by Fact 3.34.
For any type-definable with , the group is type-definable with and . If the index is , we have some with , and , so , so — a contradiction. So , and .
(3) Let be the left stabilizer of . Then is type-definable by a symmetric version of Fact 3.34. By (1), is left -invariant, so we have , and so is left -invariant. By a symmetric argument as in (2), we conclude . ∎
Example 3.40.
If is a fim type-definable subgroup of , witnessed by a -invariant fim measure , then is obviously idempotent and generically transitive.
Analogously to the case of types (Section 2.4), the following is our main question in the case of measures:
Problem 3.41.
Assume that is fim and idempotent. Is it true that then is generically transitive? Assuming is NIP?
By symmetric versions of the above considerations, one easily gets:
Remark 3.42.
Section 3.7 remains valid if one swaps “left” with “right” (including left and right stabilizers) and replaces by .
3.8. Idempotent fim measures in abelian groups
Finally, we have all of the ingredients to adapt the proof in Section 2.5 from types to measures and give a positive answer to Problem 3.41 for abelian groups in arbitrary theories. Let and be as in the previous section.
First note the idempotency of a fim measure can be iterated along a Morley sequence in :
Lemma 3.43.
Let and suppose that is fim and idempotent. Then for any formula and we have
Proof.
By induction on , the base case is the assumption on . So assume that for any and we have
(a) |
Fix , and choose a small such that is -invariant and contains the parameters of . Let and . For any (and any ) we have
(b) |
Then, using that is fim and fim measures commute with Borel definable measures,
Lemma 3.44.
Assume that is fim and idempotent. Let be arbitrary and let be the definable map
where “” is viewed as a globally defined function whose restriction to defines the group operation (see Section 2.2). Let . Then:
-
(1)
;
-
(2)
if is abelian, then for every .
Proof.
(1) By definition of .
(2) Let be arbitrary, and let be a small model containing its parameters, and so that is -invariant. We let
Using that is abelian and Lemma 3.43, for any and we have (renaming the variables when necessary):
Hence, since is abelian, is fim, and fim measures commute with Borel definable measures, we get
Finally, using results about the randomization (Theorem 3.13) and Lemma 3.44, we can show generic transitivity in the abelian case:
Theorem 3.45.
Assume that is an abelian type-definable group and is fim and idempotent. Then is generically transitive.
Proof.
By Proposition 3.37, it suffices to show that . Assume not, say
for some and some . Let be a small model containing the parameters of , and so that is invariant over . Let be as given by the moreover part of Theorem 3.13 for . Fix any , and consider the definable map given by . Then induces a continuous map from to , where and we let be defined by . That is, for every and every we have
Then and from Lemma 3.44, so by Lemma 3.44 we then have for all — contradicting the choice of . ∎
3.9. Support transitivity of idempotent measures
A tempting strategy for generalizing the arguments in Sections 2.7–2.10 with ranks from idempotent types to idempotent measures in is to apply (a continuous logic version of) the proof for types in the randomization of , assuming that the randomization preserves the corresponding property. E.g, stability is preserved [BYK09], and (real) rosiness is known to be preserved in some special cases [AGK19] (e.g. when is -minimal). We note that simplicity of is not preserved, still one gets that is NSOP1 assuming that is simple [BYCR24]. When attempting to implement this strategy, one arrives at the following natural condition connecting the behavior of measures and types in their support:
Definition 3.46.
Assume that is a type-definable group and . We say that is support transitive if for every .
Remark 3.47.
-
(1)
If is a generically stable idempotent type, then it is obviously support transitive (viewed as a Keisler measure).
-
(2)
If is generically transitive, then it is support transitive.
Indeed, note that if , , is -invariant for a small model , and , then
where the last equality follows as by assumption, and is -type-definable by Fact 3.34.
Thus, we view the following as an intermediate (and trivial in the case of types) version of our main Problem 3.41:
Problem 3.48.
Assume that is a type-definable group and is fim and idempotent. Is support transitive?
The following example (based on [CG23, Example 4.5]) illustrates that the fim assumption in Problem 3.48 cannot be relaxed to either definable or Borel-definable and finitely satisfiable, even in abelian NIP groups:
Example 3.49.
Consider , , and and , where are the unique complete -types satisfying:
-
•
,
-
•
,
-
•
,
-
•
.
Then is finitely satisfiable in (hence also Borel-invariant over by NIP) and is definable over , but neither is fim. The following are easy to verify directly:
-
(1)
, and — hence is idempotent, finitely satisfiable in , but not support-invariant;
-
(2)
likewise, , , and — hence is idempotent, definable over , but not support invariant.
Support transitivity is closely related to the algebraic properties of the semigroup induced by on the support of an idempotent measure, studied in [CG22, Section 4].
Fact 3.50.
[CG22, Corollary 4.4] Assume that is fim and idempotent. Then is a compact Hausdorff semigroup which is left-continuous, i.e. the map is continuous for each fixed .
Proposition 3.51.
Assume that is fim and idempotent. Then the following are equivalent:
-
(1)
is support transitive, i.e. for all ;
-
(2)
for any there exists such that ;
-
(3)
, where is a minimal (closed) left ideal of .
Proof.
(2) (3). By [CG22, Remark 4.17].
(3) (1). By [CG22, Corollary 4.16].
(1) (2). Let be given. By Fact 3.50, the map defined via is continuous. We will show that it has a dense image. Then by compactness of and continuity of , the image of is also closed, hence is surjective — proving the claim.
Indeed, fix some formula such that and choose a small such that is -invariant. By (1), . Let , then . Hence there exists some such that . But then by definition , hence . As was arbitrary, this shows that the image of is dense. ∎
We know that this property holds in specific examples like the circle group (e.g., see [CG22, Example 4.2]).
3.10. Idempotent measures in stable theories, revisited
It is shown in [CG22, Theorem 5.8] that every idempotent measure on a type-definable group in a stable theory is generically transitive. The proof consists of two ingredients: an analysis of the convolution semigroup on the support of an idempotent Keisler measure, and an application of a variant of Hrushovski’s group chunk theorem for partial types due to Newelski [New91].
In this section we provide an alternative argument, implementing the strategy outlined at the beginning of Section 3.9 of working in the randomization. This replaces the use of Newelski’s theorem by a direct generalization of the proof for types in stable theories from Section 2.7, and the only fact about the supports of idempotent measures that we will need is that they are support transitive.
To simplify the notation, in this section we will assume that is an -theory expanding a group. We first recall the basic results about local ranks in continuous logic, from [BY10, BYU10]. The following facts are proved under more general hypothesis in Sections 7 and 8 of [BYU10].
Fact 3.52.
Suppose that is a continuous stable theory. Let .
-
(1)
For any there exists a unique -definable extension .
-
(2)
For every and every partitioned -formula , there exists a rank function , which we call the -Cantor-Bendixson rank. More specifically, for any subset and , is an ordinal.
-
(3)
For any and such that , is the unique definable extension of if and only if for every , .
The following proposition is a standard exercise from the previous fact.
Proposition 3.53.
Let be a continuous stable theory expanding a group. Let be a monster model of and a small submodel. For any partitioned -formula we let . Then:
-
(1)
for any and , ;
-
(2)
for any and such that , we have that is the unique definable extension of if and only if for every partitioned -formula and for all , we have .
Proof.
-
(1)
Given , the map defined via is a bijective isometry. In other words, it is an automorphism of as a topometric space, and computing the rank is unaffected.
-
(2)
Follows directly from (3) of Fact 3.52. ∎
We refer to Section 3.4 for notation regarding Keisler randomizations.
Fact 3.54.
[BYK09, Theorem 5.14] If is stable, then its Keisler randomization is stable.
Proposition 3.55.
Suppose that is stable, is idempotent and support transitive. Then is generically transitive.
Proof.
Let be a bigger monster model of . Fix an atomless probability algebra and consider the randomizations , where is a monster model of . Given (note that is definably by stability), we let be as defined in Fact 3.17. Similarly, given , we let be as defined in Fact 3.17, but with respect to in place of .
Let now be idempotent and support transitive. Since is stable, there is some small model such that is -definable. Let be the unique -definable extension of . To show that is generically transitive, it suffices to prove that for every and such that , we have that . We let . By construction, and is -definable. Since is stable (Fact 3.54), is the unique global definable extension of .
We claim that then . Indeed, let be an -formula and . If is a partition of for , using that is support transitive we have:
Likewise, it is straightforward to check that , where is the constant random variable (i.e., via for all ) and is the randomization of the multiplication of the group in .
Since local rank in the stable theory is translation invariant (Proposition 3.53), we conclude that is the unique definable extension of . This implies that and in turn, . This completes the proof. ∎
Remark 3.56.
We expect that this approach could be adapted for groups definable in -minimal structures (as their randomizations are known to be real rosy [AGK19]), by developing a stratified local thorn rank in continuous logic and generalizing the proof for types in rosy (discrete) first order theories from Section 2.10. When is a simple theory, the randomization is NSOP1 (but not necessarily simple) by [BYCR24]. A local rank for NSOP1 theories is proposed in [CKR23, Section 5], but a workable stratified rank is lacking at the moment. We do not pursue these directions here.
4. Topological dynamics of and in NIP groups
In this section, we will use slightly different notation from the rest of the paper, in order to preserve continuity with the earlier work and setup in [CG22, CG23]. We let be an expansion of a group, and a monster model. Throughout this section, we assume that has NIP.
It was demonstrated in [CG22, Proposition 6.4] that then the spaces of global -invariant Keisler measures, and Keisler measures which are finitely satisfiable in (denoted and , respectively) form left-continuous compact Hausdorff semigroups with respect to definable convolution (Definition 3.35). Note that is a submonoid of . By we denote the submonoid of consisting of all global types finitely satisfiable in (viewed as -measures).
In [CG23, Theorem 6.11], the first two authors described a minimal left ideal of [and ] in terms of the Haar measure on an ideal (or Ellis) group of [resp. ]. However, this required a rather specific assumption that this ideal group is a compact topological group with the topology induced from [resp. ]. In this section, we obtain the same description in the case of , but under a more natural (from the point of view of topological dynamics) assumption that the so-called -topology on some (equivalently, every) ideal group is Hausdorff (equivalently, the ideal group with the -topology is a compact topological group). In fact, the revised Newelski’s conjecture formulated by Anand Pillay and the third author in [KP23, Conjecture 5.3] predicts that the -topology is always Hausdorff under NIP. In Section 5, we confirm this conjecture in the case when is countable, which is an important result by its own rights. In particular, in the case when is countable, our description of a minimal left ideal of does not require any assumption on the ideal group.
As discussed in the introduction, the -topology plays an essential role in many important structural results in abstract topological dynamics, including the recent theorem of Glasner on the structure of tame, metrizable, minimal flows [Gla18]. In fact, our proof of the revised Newelski’s conjecture for countable will be deduced using this theorem of Glasner.
The reason why our proof of the revised Newelski’s conjecture requires the countability of assumption is to guarantee that certain flows of types are metrizable in order to be able to apply the aforementioned theorem of Glasner. The reason why we focus only on and (and not on and ) is that is isomorphic to the Ellis semigroup of the -flow and so we have the -topology on the ideal group of at our disposal. For the revised Newelski’s conjecture we will also use a well-known general principle that NIP implies tameness for various flows of types [CS18, Iba16, KR20].
4.1. Preliminaries from topological dynamics
Definition 4.1.
A -flow is a pair , where is an abstract group acting (on the left) by homeomorphisms on a compact Hausdorff space .
Definition 4.2.
If is a flow, then its Ellis semigroup, denoted by or just , is the (pointwise) closure in of the set of functions for .
Fact 4.3.
(see e.g. [Aus88]) If is a flow, then is a compact left topological semigroup (i.e. it is a semigroup with the composition as its semigroup operation, and the composition is continuous on the left, i.e. for any the map is continuous). It is also a -flow with .
The next fact is folklore. Thanks to this fact is always identified with .
Fact 4.4.
The function given by , where is defined by , is an isomorphism of semigroups and -flows.
The following is a fundamental theorem of Ellis on the basic structure of Ellis semigroups (see e.g. [Ell69, Corollary 2.10 and Propositions 3.5 and 3.6] or Proposition 2.3 of [Gla76, Section I.2]). We will use it freely without an explicit reference.
Fact 4.5 (Ellis’ Theorem).
Suppose is a compact Hausdorff left topological semigroup (e.g. the enveloping semigroup of a flow). Then has a minimal left ideal . Furthermore, for any such ideal :
-
(1)
is closed;
-
(2)
for any element and idempotent we have , and ;
-
(3)
, where ranges over all idempotents in ; in particular, contains an idempotent;
-
(4)
for any idempotent , the set is a subgroup of with the neutral element .
Moreover, all the groups (where ranges over all minimal left ideals and over all idempotents in ) are isomorphic. In the model theory literature, the isomorphism type of all these groups (or any of these groups) is called the ideal (or Ellis) group of ; if , we call this group the ideal (or Ellis) group of the flow .
We will use the following fact, which gives us an explicit isomorphism from Fact 4.5 between any two ideal groups in a given minimal left ideal. The context is as in Fact 4.5.
Fact 4.6.
If and are idempotents in , then defines an isomorphism .
Proof.
The map is a homomorphism, because . In the same way, given by is a homomorphism. And it is clear that is the inverse of . ∎
We will also need the following observation which was Lemma 3.5 in the first arXiv version of [KLM22] (the section with this results was removed in the published version).
Lemma 4.7.
Let be a compact left topological semigroup, a minimal left ideal of , and an idempotent. Then the closure of is a (disjoint) union of ideal groups. In particular, is a subsemigroup of .
Proof.
Note that for every , is an ideal group. Namely, for some idempotent . Thus, , but also , because , where is the inverse of in the ideal group .
Let now (as is closed). By the first paragraph, it suffices to prove that . Since , we have . Take any . Then for some . Since , we have that is the limit point of a net . By left continuity, , so as for all ’s. ∎
Most of the statements in the next fact are contained in [Gla76, Section IX.1]. There, the author considers the special case of and defines in a slightly different (but equivalent) way. However, as pointed out in [KP17, Section 2] and [KPR18, Section 1.1], many of the proofs from [Gla76, Section IX.1] go through in the general context. A very nice exposition of this material (with all the proofs) in the general context can be found in Appendix A of [Rze18].
Fact 4.8 (The -topology on the ideal group in an Ellis semigroup).
Consider the Ellis semigroup of a flow , let be a minimal left ideal of and an idempotent.
-
(1)
For each , , we write for the set of all limits of nets , where are such that and .
-
(2)
The formula defines a closure operator on . It can also be (equivalently) defined as . We call the topology on induced by this operator the -topology.
-
(3)
If (a net in ) converges to (the closure of in ), then converges to in the -topology.
-
(4)
The -topology on coarsens the subspace topology inherited from .
-
(5)
with the -topology is a quasi-compact, semitopological group (that is, the group operation is separately continuous) in which the inversion is continuous.
-
(6)
All the groups (where ranges over all minimal left ideals of and over all idempotents in ) equipped with the -topology are isomorphic as semitopological groups. In particular, the map from Fact 4.6 is a topological isomorphism.
Corollary 4.9.
If the -topology on is Hausdorff, then is a compact topological group.
The following result is Lemma 3.1 in [KPR18].
Fact 4.10.
Let be a flow. Let be a minimal left ideal of and an idempotent. Then the function (where is the closure of in the topology of ) defined by the formula has the property that for any continuous function , where is a regular topological space and is equipped with the -topology, the composition is continuous, where is equipped with subspace topology from . In particular, if is Hausdorff with the -topology, then is continuous.
In the model-theoretic context of the -flow (with the action of by left translations), the following fact is folklore (see e.g. [New12, Pil13]).
Fact 4.11.
The function given by , where is defined by , is an isomorphism of semigroups and -flows.
Thanks to this fact, Fact 4.8 can be applied directly to in place of , which we do without further explanations.
Note, however, that Fact 4.11 does not hold for in place of , because the -orbit of need not be dense in .
4.2. Minimal left ideal of
Recall that is an expansion of a group. Fix which is -saturated. Recall that in this section we assume that is NIP.
Definition 4.12.
For a formula and , define a new formula
In the next fact, stands for . By the proof of [HP11, Proposition 2.6] (where, if is of the form for some and , is chosen depending on ), we have:
Fact 4.13.
Let and be any formula. Let . Then, there exists a positive such that , where denotes the complement of a set and
The following lemma is the key new step needed to adapt the arguments from [CG23, Section 6.2] to our general setting here. From now on, let be a minimal left ideal in and an idempotent. For , we will typically write instead of .
Lemma 4.14.
Let be any formula. Assume that the -topology on the ideal group is Hausdorff. Then the subset of is constructible, and so Borel (in the -topology).
Proof.
By Fact 4.10 and the assumption that is Hausdorff, the function (where is the closure of in the topology of ) given by is continuous. Note that is a subsemigroup of which coincides with the set of all idempotents in . This follows from Lemma 4.7 and the fact that for every the restriction is a group isomorphism by Fact 4.6.
Put
Let be a monster model in which is small. Let be the unique extension of to a type in which is finitely satisfiable in . Pick (in a yet bigger monster model). Put
Then .
Take , , and from Fact 4.13 applied for in place of and for . Then . Define:
Note that all the sets are contained in , as is a semigroup by Lemma 4.7.
Claim.
-
(1)
.
-
(2)
.
-
(3)
, , and .
-
(4)
and .
-
(5)
.
-
(6)
and are closed.
-
(7)
is constructible.
Proof.
(1) Let be given by . Note that all of the sets are -invariant (more precisely, invariant over and the parameters of in ), hence they are unions of sets of realizations in of complete types over . Then we have
So .
(2) Take and . Then , which by definition implies .
Take . Then, by Lemma 4.7, for some (where is the set of all idempotents in ). Then and , so .
(3) The first two equalities follow from the definition of and and the fact that satisfies (as for ). To show in the third equality, take any and . Then by the first equality. Moreover, , as otherwise , so by the second equality, where is such that , a contradiction. To see , take any . We have that for some . So .
(4) follows as on the last line of (3).
(5) Take . Then for some . Hence, for some . Since , by (3), we get that , so, by (4), . Thus, which is contained in by (1). Hence, by (2), we conclude that .
Suppose for a contradiction that there is some . Using (1) and (2), let be maximal for which there is such that . Then by (1) and (3). So . This together with and (4) implies that , so for some . Then , so for some . Choose with . Then , and so which is contained in by (2) (as ). Therefore, by (1), for some . Since , we get . As , we get a contradiction with the maximality of .
(6) follows as is continuous by Fact 4.10 (this is the only place in the proof where we use the assumption that is Hausdorff) and are closed.
(7) To show the first equality, it is enough to prove that and are unions of fibers of . Let us prove it for ; the case of is the same. So consider any and with . We have and for some . Then and . So, since is an isomorphism, we get that by (3).
The second equality follows since:
Hence, is constructible by (6). ∎
By item (5) of the claim, we get , which is a constructible set by item (7) of the claim. ∎
Proposition 4.15.
Using this, the material from Lemma 6.9 to Corollary 6.12 of [CG23] goes through word for word. In particular, we get the following lemma and the main theorem describing a minimal left ideal of the semigroup , under a more natural assumption than the property CIG1 (requiring that is compact with respect to the induced topology instead of the -topology) assumed in [CG23].
Lemma 4.16.
Assume that the -topology on is Hausdorff. Then for all , where is the Dirac measure at .
Let .
Theorem 4.17.
Assume that the -topology on is Hausdorff. Then is a minimal left ideal of , and is an idempotent which belongs to .
In the next section we will see that the assumption that the -topology on is Hausdorff is always satisfied when is a countable NIP group.
5. Revised Newelski’s conjecture for countable NIP groups
The goal of this section is to prove the revised Newelski’s conjecture (see [KP23, Conjecture 5.3]) working over a countable model. We consider here a standard context for this conjecture (originally from [New09, Section 4], but see also [CPS14]) which is slightly more general than in the previous Section 4. Let be a model of a NIP theory, a 0-definable group in , and an -saturated elementary extension. By we denote the space of all complete external types over concentrated on , i.e. the space of ultrafilters of externally definable subsets of . It is a -flow with the action given by left translation, which is naturally isomorphic as a -flow to the -flow of all complete types over concentrated on and finitely satisfiable in . Note that the previous context of Section 4 is a special case when and . On we have the left continuous semigroup operation defined in the same way as for (i.e. for any , such that is finitely satisfiable in ), and Fact 4.11 still holds for it.
The following revised Newelski’s conjecture was stated in [KP23, Conjecture 5.3].
Conjecture 5.1.
Assume that is NIP. Let be a minimal left ideal of and an idempotent. Then the -topology on is Hausdorff.
A background around this conjecture, including an explanation that it is a weakening of Newelski’s conjecture, is given in the introduction, and in more details in the long paragraph preceding Conjecture 5.3 in [KP23] and short paragraph following it.
In order to prove Conjecture 5.1 for countable , first we will deduce from the main theorem of [Gla18] on the structure of tame, metrizable, minimal flows that each such flow has Hausdorff ideal group. Then Conjecture 5.1 for countable will follow using this fact and a presentation of as an inverse limit of certain metrizable flows. The topological dynamical material below is rather standard, but it requires recalling quite a few notions and basic facts about them, and making some observations.
From now on, until we say otherwise, we are in the general abstract context of -flows and homomorphisms between them, where is an arbitrary abstract (not necessarily definable in a NIP theory) group. We let , , etc. be -flows, which we will sometimes denote simply as , etc.
For a proof of the following fact see [Rze18, Proposition 5.41].
Fact 5.2.
Let be an epimorphism of -flows. Then given by is a well-defined semigroup and -flow epimorphism. If is a minimal left ideal of and an idempotent, then is a minimal left ideal of and is an idempotent in . Moreover, is a group epimorphism and topological quotient map with respect to the -topologies.
Remark 5.3.
If and are flows for which there exists a semigroup and flow epimorphism , then it is unique. In particular, the epimorphism in Fact 5.2 does not depend on the choice of the epimorphism .
Proof.
Since is a semigroup epimorphism, it satisfies . Let be the left translation by , and similarly . Then . By definition, is the closure of , so is unique (as it is continuous). ∎
Let be the Stone-Čech compactification of and the principal ultrafilter at . As e.g. explained on page 9 of [Gla76], the -ambit (where by a -ambit we mean a -flow with a distinguished point with dense orbit) is universal, and so there is a unique left continuous semigroup operation on extending the action of by left translation. (In fact, it is precisely the operation on for expanded by predicates for all subsets of ). For any flow , universality of also yields a unique action of the semigroup on which is left-continuous and extends the action of , and which we will denote by . Fix a minimal left ideal of and an idempotent . Using this action, for any -flow with a distinguished point such that (note that such an always exists), the Galois group of is defined as:
it is a -closed subgroup of (see [Gla76, Page 13]). In the topological dynamics literature, this group is sometimes called the Ellis group of , e.g. see [Gla76, Page 13], where it is denoted , for its basic properties.
There is an obvious semigroup and -flow epimorphism given by . It is unique by Remark 5.3. As in Fact 4.11, is naturally isomorphic to via , where . Using this identification, for any -flow epimorphism , the induced map from Fact 5.2 coincides with .
Remark 5.4.
Let and be as above. Let be a flow and the unique epimorphism defined above. Let and .
-
(1)
For every with , .
-
(2)
For every , .
-
(3)
.
-
(4)
is a group epimorphism and topological quotient map with respect to the -topologies.
Proof.
It is clear that is a minimal left ideal of , an idempotent, and a group epimorphism.
(1) Take , i.e. . Then . Hence, .
(2) Take , i.e. for some . Then .
(3) The inclusion follows from (1) and (2). For the opposite inclusion, consider any . In order to show that , it is enough to check that (because for any we have , and so the map given by is injective, in fact a group monomorphism, and as is an idempotent). But this is trivial by the choice of : .
(4) In the situation when has a dense orbit (which is for example the case when is minimal), this follows from Fact 5.2, the existence of an epimorphism (as is a universal -ambit), and the observation that made just before Remark 5.4. In general, it follows from the straightforward generalization of Fact 5.2 stated in [KLM22, Fact 2.3]. ∎
Definition 5.5.
A -flow epimorphism is almost 1-1 if the set is dense in .
Remark 5.6.
If is minimal and is almost 1-1, then is also minimal.
Proof.
We will show that for every we have . This implies that for every (because is dense in and is closed in ), which means that is minimal.
So fix , and consider any . Since is minimal, we can find such that . Pick satisfying . Then . Since , we conclude that . ∎
Lemma 5.7.
If is minimal and is almost 1-1, then the group homomorphism is a topological isomorphism (in the -topologies), where is a minimal left ideal of and an idempotent.
Proof.
By Remark 5.6, is minimal, so for every . Pick ; then there is such that . Choose an idempotent so that . Then , so , and hence by idempotence of .
Since the diagram
commutes (by definition of ) and, by Fact 4.8(6), the horizontal arrows are isomorphisms of semitopological groups, it is enough to show that is an isomorphism of semitopological groups. By Fact 5.2, it is a group epimorphism and topological quotient map, so it remains to show that it is injective.
Suppose for a contradiction that is non-trivial, i.e. there is such that . Then , so . On the other hand, by the first sentence of the proof, . So there is with , and hence (where is the inverse of computed in ). As , and is a group morphism, we get (where the last equality follows from the first paragraph) — a contradiction. ∎
A pair of points of a flow is called proximal if there is such that ; it is called distal if or is not proximal. Let denote the collection of all proximal pairs of points in . The flow is said to be proximal when , and distal when .
Fact 5.8.
The ideal group of every proximal flow is trivial.
Proof.
Let be a minimal left ideal of and an idempotent. From [Gla76, Chapter I, Proposition 3.2(3)], it follows that each pair of points in is distal, so, by proximality, is a singleton, say . Then for any in , say with , and any , we have . So , hence . ∎
Whenever is a -flow epimorphism, let
Definition 5.9.
A -flow epimorphism is said to be:
-
(1)
equicontinuous (or almost periodic) if for every which is an open neighborhood of the diagonal , there exists a neighborhood of such that for every ;
-
(2)
distal if .
It is sometimes assumed that the flows in the definition of equicontinuous extensions are minimal. On page 100 of [Gla76], Glasner defines almost periodic extensions of minimal flows in a different way. A proof that both definitions are equivalent for minimal flows can be found in [Aus88, Chapter 14, Theorem 1].
The following remark is well-known and follows from an argument on page 4 of [Gla76].
Remark 5.10.
An equicontinuous epimorphism of flows is distal.
Proposition 5.11.
If is an equicontinuous epimorphism of minimal -flows and has a trivial ideal group, then the ideal group of is Hausdorff (with respect to the -topology).
Proof.
Choose a minimal left ideal in , an idempotent , and an element with . Let be the unique semigroup and -flow epimorphism considered before and in Remark 5.4. Put and . So is the ideal group of . Put . Then (as , where is a net in converging to in ). Let and . So is the ideal group of which is trivial by assumption.
Clearly is the unique semigroup and -flow epimorphism from to , and is a group epimorphism. Hence, as is trivial, . As by Remark 5.4(1) , we conclude that .
Let be a -closed subgroup of , and let
As is almost periodic, [Gla76, Chapter IX, Theorem 2.1(4)] yields , and together with the conclusion of the last paragraph this implies . Note that we proved it for any with , in particular for any (by Remark 5.4(2)). On the other hand, by Remark 5.4(3), . Hence, .
There are many equivalent definitions of tame flows (see Theorems 2.4, 3.2 and Definition 3.1 in [GM18]). We give the one which immediately points to a strong connection with the NIP property in model theory.
A sequence of real valued functions on a set is said to be independent if there exist real numbers such that
for all finite disjoint subsets of .
Definition 5.12.
Let be a flow. A function (i.e. a continuous real valued function on ) is tame if the the family of translates does not contain an infinite independent sequence (where ). The flow is tame if all functions in are tame.
The following fact is a part of the information contained in the main theorem (Theorem 5.3) of [Gla18] on the structure of tame, metrizable, minimal flows.
Fact 5.13.
Let be a tame, metrizable, minimal flow. Then there exists the following commutative diagram of -flow epimorphisms
where:
-
(1)
is minimal;
-
(2)
is proximal;
-
(3)
are almost 1-1;
-
(4)
is equicontinuous.
We discussed all of the notions and facts above in order to deduce the following corollary.
Corollary 5.14.
The -topology on the ideal group of any tame, metrizable, minimal flow is Hausdorff.
Proof.
We will be referring to items (1)–(4) in Fact 5.13. By (1), (3) and Remark 5.6, is minimal, and so are , , and as homomorphic images of . By (2) and Fact 5.8, the ideal group of is trivial, and so is the ideal group of by (3) and Lemma 5.7. Hence, using (4) and Proposition 5.11, we get that the ideal group of is Hausdorff, and so is the ideal group of by (3) and Lemma 5.7. Therefore, the ideal groups of and are both Hausdorff by Fact 5.2, because they are quotients of a compact topological group (namely, the ideal group of ) by closed subgroups. ∎
To apply this general corollary to our model-theoretic context, we need one more general observation, namely Lemma 5.16. To prove it, we have to recall a description of the -closure that was stated as Lemma 3.11 in the first arXiv version of [KLM22] (the relevant section of [KLM22] was removed in the published version).
Fact 5.15.
Let be a flow, a minimal left ideal of , and an idempotent. Then for every , the -closure can be described as the set of all limits contained in of nets such that , , and .
Proof.
Consider . Then, by the definition of the -topology, there are nets and such that and . Note that , as . Put for all . By left continuity, we have that . Furthermore, .
Conversely, consider any for which there are nets and such that and . Since each can be approximated by elements of and the semigroup operation is left continuous, one can find a subnet of and a net such that and , which means that . ∎
Lemma 5.16.
Let be a flow, a minimal left ideal in . Then there exists a minimal left ideal of and a semigroup and -flow isomorphism from to . In particular, the ideal groups of the -flows and are isomorphic as semitopological groups (with the -topologies).
Proof.
Denote the semigroup operation on by and on by (although both are compositions of functions, but on different sets). Let be given by , where . It is easy to check that is a semigroup and -flow monomorphism. Put . Thus, is a semigroup and -flow isomorphism. We need to check that is a minimal left ideal in . For that, first note that is a minimal left ideal in itself. Indeed, if is a left ideal in , then for any and , taking an idempotent such that , we have , and so as ( is a left ideal in ). Hence, is a left ideal in which is contained in , so it must be equal to by minimality of . Since is a minimal left ideal in itself, we get that
-
(1)
is a minimal left ideal in itself.
On the other hand,
-
(2)
is a left ideal in .
To prove (2), consider any and , and we need to show that . Using limits of nets, we easily see that there is such that . For any we have: , where satisfies for all (such exists by Fact 4.4). Hence, (and as is a left ideal).
By (1) and (2), we get that is a minimal left ideal of .
Thus, is the ideal group of , and is a group isomorphism, where is an idempotent. This isomorphism is topological (with respect to the -topologies) by Fact 5.15 (which expresses the -closure in terms of the semigroup operation and convergence within the minimal left ideal in question) and the above observation that is an isomorphism of left topological semigroups. ∎
We can finally prove Conjecture 5.1 for countable .
Theorem 5.17.
Let be a countable model of a theory with NIP and be -saturated. Let be a minimal left ideal of and an idempotent. Then the -topology on is Hausdorff.
Proof.
Let be the Shelah expansion of obtained by adding predicates for all externally definable subsets of for all . By Shelah’s theorem [She09] (see also [CS13]), we know that has quantifier elimination, NIP, and all types in are definable (i.e. all externally definable subsets of are definable). It follows that the Boolean algebra of externally definable subset of with respect to the original language coincides with the Boolean algebra of definable subsets of in the sense of the expanded language. Hence, . Thus, without loss of generality, we may assume that is a countable model of an NIP theory such that all types in are definable. Then , and the semigroup operation on is given by , where , , and is finitely satisfiable in .
It is well-known, and observed first time in the introduction of [CS18], that NIP implies that is a tame flow. (A standard way to see it is to note that, by NIP, the characteristic functions of all the clopens in are tame (in the sense of Definition 5.12) and separate points, and so, by Stone-Weierstrass theorem, they generate a dense subalgebra of ; then use the fact that tame functions on form a closed subalgebra of to conclude that all functions in are tame.) However, is neither metrizable (even when the original language of was countable, the expanded language of is usually uncountable) nor minimal, so we cannot apply Corollary 5.14 directly to .
Let range over all finite collections of definable subsets of . For any such , let be the Boolean -algebra (so closed under left translations by the elements of ) of subsets of generated by , and denote by the space of all ultrafilters of . Note that is naturally a -flow (the action is by left translations), and as -flows. Let be any minimal left ideal (and so minimal subflow) of . Let be the image of under the restriction map. It is clearly a minimal subflow of , and the above isomorphism induces a -flow isomorphism (see [Rze18, Lemma 6.42]).
Since is tame, so is as a quotient of , and so is as a subflow of (using Definition 5.12 and Tietze’s extension theorem, or see e.g. [KR20, Fact 4.20]). Moreover, is metrizable since is countable by finiteness of and countability of (and this is the only place where we the assumption that is countable). Hence, is metrizable.
Summarizing, is a tame, metrizable, minimal flow, and hence its ideal group is Hausdorff by Corollary 5.14.
On the other hand, since , [Rze18, Lemma 6.42] implies that the ideal group of equipped with the -topology is topologically isomorphic to an inverse limit of the ideal groups (with the -topologies) of the flows .
By the last two paragraphs, we conclude that the ideal group of is Hausdorff.
Corollary 5.18.
Assume that is countable and is NIP, and let be a minimal left ideal in and an idempotent. Then is a minimal left ideal of , and is an idempotent which belongs to .
6. Acknowledgements
We thank Aaron Anderson, Martin Hils, Anand Pillay, Sergei Starchenko, Atticus Stonestrom and Mariana Vicaria for helpful conversations. Chernikov was partially supported by the NSF CAREER grant DMS-1651321 and by the NSF Research Grant DMS-2246598. Krupiński was supported by the Narodowe Centrum Nauki grant no. 2016/22/E/ST1/00450.
References
- [ACP14] Hans Adler, Enrique Casanovas and Anand Pillay “Generic stability and stability” In The Journal of Symbolic Logic JSTOR, 2014, pp. 179–185
- [Adl07] Hans Adler “Strong theories, burden, and weight” In preprint 2008, 2007
- [AGK19] Uri Andrews, Isaac Goldbring and H Jerome Keisler “Independence in randomizations” In Journal of Mathematical Logic 19.01 World Scientific, 2019, pp. 1950005
- [And23] Aaron Anderson “Generically Stable Measures and Distal Regularity in Continuous Logic” In Preprint, arXiv:2310.06787, 2023
- [Aus88] Joseph Auslander “Minimal Flows and Their Extensions”, Mathematics Studies 153 Amsterdam: North-Holland, 1988
- [Bay18] Martin Bays “Geometric stability theory” In Lectures in model theory, 2018, pp. 29–58
- [BMPW16] Thomas Blossier, Amador Martin-Pizarro and Frank O Wagner “A la recherche du tore perdu” In The Journal of Symbolic Logic 81.1 Cambridge University Press, 2016, pp. 1–31
- [BY09] Itaï Ben Yaacov “Transfer of properties between measures and random types”, Unpublished research note, http://math.univ-lyon1.fr/~begnac/articles/MsrPrps.pdf, 2009
- [BY10] Itaï Ben Yaacov “Stability and stable groups in continuous logic” In The Journal of Symbolic Logic 75.3 Cambridge University Press, 2010, pp. 1111–1136
- [BYCR24] Itaï Ben Yaacov, Artem Chernikov and Nick Ramsey “Randomization and Keisler measures in NSOP1 theories” In Preprint, 2024
- [BYK09] Itaï Ben Yaacov and Jerome H Keisler “Randomizations of models as metric structures” In Confluentes Mathematici 1.02 World Scientific, 2009, pp. 197–223
- [BYU10] Itaï Ben Yaacov and Alexander Usvyatsov “Continuous first order logic and local stability” In Transactions of the American Mathematical Society 362.10, 2010, pp. 5213–5259
- [Cas11] Enrique Casanovas “More on NIP and related topics (Lecture notes of the Model Theory Seminar, University of Barcelona)”, http://www.ub.edu/modeltheory/documentos/nip2.pdf, 2011
- [CG20] Gabriel Conant and Kyle Gannon “Remarks on generic stability in independent theories” In Annals of Pure and Applied Logic 171.2 Elsevier, 2020, pp. 102736
- [CG22] Artem Chernikov and Kyle Gannon “Definable convolution and idempotent Keisler measures” In Israel Journal of Mathematics 248.1 Springer, 2022, pp. 271–314
- [CG23] Artem Chernikov and Kyle Gannon “Definable convolution and idempotent Keisler measures, II” In Model Theory 2.2 Mathematical Sciences Publishers, 2023, pp. 185–232
- [CGH23] Gabriel Conant, Kyle Gannon and James Hanson “Generic stability, randomizations, and NIP formulas” In Preprint, arXiv:2308.01801, 2023
- [CGH23a] Gabriel Conant, Kyle Gannon and James Hanson “Keisler measures in the wild” In Model Theory 2.1 Mathematical Sciences Publishers, 2023, pp. 1–67
- [Che14] Artem Chernikov “Theories without the tree property of the second kind” In Annals of Pure and Applied Logic 165.2 Elsevier, 2014, pp. 695–723
- [Che18] Artem Chernikov “Model theory, Keisler measures, and groups” In Bulletin of Symbolic Logic 24.3 Cambridge University Press, 2018, pp. 336–339
- [CK12] Artem Chernikov and Itay Kaplan “Forking and dividing in NTP2 theories” In The Journal of Symbolic Logic 77.1 Cambridge University Press, 2012, pp. 1–20
- [CKR23] Artem Chernikov, Byunghan Kim and Nicholas Ramsey “Transitivity, Lowness, and Ranks in NSOP1 Theories” In The Journal of Symbolic Logic 88.3, 2023, pp. 919–946
- [Coh60] Paul J Cohen “On a conjecture of Littlewood and idempotent measures” In American Journal of Mathematics 82.2 JSTOR, 1960, pp. 191–212
- [CP12] Annalisa Conversano and Anand Pillay “Connected components of definable groups and o-minimality I” In Advances in Mathematics 231.2 Elsevier, 2012, pp. 605–623
- [CPS14] Artem Chernikov, Anand Pillay and Pierre Simon “External definability and groups in NIP theories” In Journal of the London Mathematical Society 90.1 Oxford University Press, 2014, pp. 213–240
- [CS13] Artem Chernikov and Pierre Simon “Externally definable sets and dependent pairs” In Israel Journal of Mathematics 194.1 Springer, 2013, pp. 409–425
- [CS18] Artem Chernikov and Pierre Simon “Definably amenable NIP groups” In J. Amer. Math. Soc. 31.3, 2018, pp. 609–641
- [CS21] Artem Chernikov and Sergei Starchenko “Definable regularity lemmas for NIP hypergraphs” In The Quarterly Journal of Mathematics 72.4 Oxford University Press UK, 2021, pp. 1401–1433
- [DK12] Jan Dobrowolski and Krzysztof Krupiński “On -categorical, generically stable groups” In The Journal of Symbolic Logic 77.3 Cambridge University Press, 2012, pp. 1047–1056
- [EKP08] Clifton Ealy, Krzysztof Krupiński and Anand Pillay “Superrosy dependent groups having finitely satisfiable generics” In Ann. Pure Appl. Logic 151.1 Elsevier BV, 2008, pp. 1–21
- [Ell57] Robert Ellis “Locally compact transformation groups” In Duke Mathematical Journal 24.2 Duke University Press, 1957, pp. 119 –125
- [Ell69] Robert Ellis “Lectures on topological dynamics”, Mathematics lecture note series 28 W.A. Benjamin, Inc., 1969
- [GK15] Jakub Gismatullin and Krzysztof Krupiński “On model-theoretic connected components in some group extensions” In Journal of Mathematical Logic 15.02 World Scientific, 2015, pp. 1550009
- [Gla07] Eli Glasner “The structure of tame minimal dynamical systems” In Ergodic Theory and Dynamical Systems 27.6 Cambridge University Press, 2007, pp. 1819–1837
- [Gla18] Eli Glasner “The structure of tame minimal dynamical systems for general groups” In Invent. Math. 211, 2018, pp. 213–244
- [Gla76] Shmuel Glasner “Proximal Flows”, Lecture Notes in Mathematics 517 Berlin: Springer-Verlag, 1976
- [Gli59] Irving Glicksberg “Convolution semigroups of measures.” In Pacific Journal of Mathematics 9.1 Pacific Journal of Mathematics, 1959, pp. 51–67
- [GM18] Eli Glasner and Michael Megrelishvili “More on tame dynamical systems” In Ergodic Theory and Dynamical Systems in their Interactions with Arithmetics and Combinatorics: CIRM Jean-Morlet Chair, Fall 2016 Springer, 2018, pp. 351–392
- [GOU13] Darío García, Alf Onshuus and Alexander Usvyatsov “Generic stability, forking, and þ-forking” In Transactions of the American Mathematical Society 365.1, 2013, pp. 1–22
- [GPP15] Jakub Gismatullin, Davide Penazzi and Anand Pillay “Some model theory of ” In Fundamenta Mathematicae 229.2 Institute of Mathematics Polish Academy of Sciences, 2015, pp. 117–128
- [HO10] Assaf Hasson and Alf Onshuus “Stable types in rosy theories” In The Journal of Symbolic Logic 75.4 Cambridge University Press, 2010, pp. 1211–1230
- [HP11] Ehud Hrushovski and Anand Pillay “On NIP and invariant measures” In J. Eur. Math. Soc. 13.4 European Mathematical Society Publishing House, 2011, pp. 1005–1061
- [HP18] Mike Haskel and Anand Pillay “On maximal stable quotients of definable groups in NIP theories” In The Journal of Symbolic Logic 83.1 Cambridge University Press, 2018, pp. 117–122
- [HPP08] Ehud Hrushovski, Ya’acov Peterzil and Anand Pillay “Groups, measures, and the NIP” In Journal of the American Mathematical Society 21.2, 2008, pp. 563–596
- [HPS13] Ehud Hrushovski, Anand Pillay and Pierre Simon “Generically stable and smooth measures in NIP theories” In Transactions of the American Mathematical Society 365.5, 2013, pp. 2341–2366
- [HRK19] Ehud Hrushovski and Silvain Rideau-Kikuchi “Valued fields, metastable groups” In Selecta Mathematica 25.3 Springer, 2019
- [Hru12] Ehud Hrushovski “Stable group theory and approximate subgroups” In Journal of the American Mathematical Society 25.1, 2012, pp. 189–243
- [Hru19] Ehud Hrushovski “Definability patterns and their symmetries”, 2019 arXiv:1911.01129 [math.LO]
- [Hru20] Ehud Hrushovski “Beyond the Lascar Group”, 2020 arXiv:2011.12009 [math.LO]
- [Hru90] Ehud Hrushovski “Unidimensional theories are superstable” In Annals of Pure and Applied Logic 50.2 Elsevier, 1990, pp. 117–137
- [Iba16] Tomás Ibarlucía “The dynamical hierarchy for Roelcke precompact Polish groups” In Israel Journal of Mathematics 215.2 Springer, 2016, pp. 965–1009
- [Kha22] Karim Khanaki “Generic stability and modes of convergence” In Preprint, arXiv:2204.03910, 2022
- [KI40] Yukiyosi Kawada and Kiyosi Itô “On the probability distribution on a compact group. I” In Proceedings of the Physico-Mathematical Society of Japan. 3rd Series 22.12 THE PHYSICAL SOCIETY OF JAPAN, The Mathematical Society of Japan, 1940, pp. 977–998
- [KLM22] Krzysztof Krupiński, Junguk Lee and Slavko Moconja “Ramsey theory and topological dynamics for first order theories” In Trans. Amer. Math. Soc. 375.4, 2022, pp. 2553–2596
- [KP17] Krzysztof Krupiński and Anand Pillay “Generalized Bohr compactification and model-theoretic connected components” In Math. Proc. Cambridge 163.2 Cambridge University Press, 2017, pp. 219–249
- [KP23] Krzysztof Krupiński and Anand Pillay “Generalized locally compact models for approximate groups”, 2023 arXiv:2310.20683 [math.LO]
- [KPR18] Krzysztof Krupiński, Anand Pillay and Tomasz Rzepecki “Topological dynamics and the complexity of strong types” In Israel J. Math. 228, 2018, pp. 863–932
- [KR20] Krzysztof Krupiński and Tomasz Rzepecki “Galois groups as quotients of Polish groups” In J. Math. Log. 20.03, 2020, pp. 2050018
- [New09] Ludomir Newelski “Topological dynamics of definable group actions” In The Journal of Symbolic Logic 74.1 Cambridge University Press, 2009, pp. 50–72
- [New12] Ludomir Newelski “Model theoretic aspects of the Ellis semigroup” In Israel Journal of Mathematics 190.1 Springer, 2012, pp. 477–507
- [New91] Ludomir Newelski “On type definable subgroups of a stable group” In Notre Dame journal of formal logic 32.2 University of Notre Dame, 1991, pp. 173–187
- [Pil13] Anand Pillay “Topological dynamics and definable groups” In The Journal of Symbolic Logic 78.2 Cambridge University Press, 2013, pp. 657–666
- [Pil96] Anand Pillay “Geometric stability theory” Oxford University Press, 1996
- [Poi01] Bruno Poizat “Stable groups” American Mathematical Soc., 2001
- [PT11] Anand Pillay and Predrag Tanović “Generic stability, regularity, and quasiminimality. In: Models, logics, and higher-dimensional categories”, CRM proceedings & lecture notes Providence, RI: American Mathematical Society, 2011
- [Pym62] John Pym “Idempotent measures on semigroups” In Pacific Journal of Mathematics 12.2 Mathematical Sciences Publishers, 1962, pp. 685–698
- [Rud59] Walter Rudin “Idempotent measures on Abelian groups” In Pacific Journal of Mathematics 9.1 Pacific Journal of Mathematics, 1959, pp. 195–209
- [Rze18] Tomasz Rzepecki “Bounded Invariant Equivalence Relations, Ph.D. thesis, Wrocław”, 2018 arXiv:1810.05113 [math.LO]
- [She09] Saharon Shelah “Dependent first order theories, continued” In Isr. J. Math. 173, 2009, pp. 1–60
- [Sim03] Patrick Simonetta “An example of a C-minimal group which is not abelian-by-finite” In Proceedings of the American Mathematical Society 131.12, 2003, pp. 3913–3917
- [Sim15] Pierre Simon “A guide to NIP theories” Cambridge University Press, 2015
- [Sim20] Pierre Simon “On Amalgamation in NTP2 Theories and Generically Simple Generics” In Notre Dame Journal of Formal Logic 61.2, 2020
- [Sto23] Atticus Stonestrom “On non-abelian dp-minimal groups” In Preprint, arXiv:2308.04209, 2023
- [Wag00] Frank Wagner “Simple Theories” Kluwer, Dordrecht, 2000
- [Wag24] Frank O Wagner “dp-minimal groups” In Preprint, arXiv:2403.12111, 2024
- [Wan22] Paul Z Wang “The group configuration theorem for generically stable types” In Preprint, arXiv:2209.03081, 2022
- [Wen54] J. G. Wendel “Haar measure and the semigroup of measures on a compact group” In Proceedings of the American Mathematical Society 5.6 JSTOR, 1954, pp. 923–929